The Cuckoos

  • 67 1,268 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Bird Families of the World A series of authoritative, illustrated handbooks of which this is the 15th volume to be published

Series editors ROBERT B. PAYNE Chief editor MICHAEL D. SORENSON KAREN KLITZ JOHN MEGAHAN Robert Payne is an ornithologist at the University of Michigan, where he is Professor of Zoology in the Department of Ecology and Evolutionary Biology, and Curator of Birds in the Museum of Zoology. Earlier ornithological positions include two years postdoc experience at the Fitzpatrick Institute of Ornithology at the University of Cape Town, in a study of African brood-parasitic cuckoos and finches, and three years teaching at the University of Oklahoma in the Great Plains. He is interested in parental care and brood parasitism of birds, evolution of the brood-parasitic finches and cuckoos and their hosts, nestling mimicry of the parental hosts by the brood parasites, behavioral imprinting and how this may lead to the origin of new species of brood-parasitic birds, and the biology of bird song. Over time his interests have developed through a mix of travel to watch birds in the field, experimental studies to ask questions about behavior development, and systematics using the museum collections. He has about 190 publications on birds, many on brood parasitism and population biology. He works with others who work well with the molecular systematics of birds. In field research he has studied the variation of the broodparasitic finches and their host species associations in Africa, and after this book is completed he plans to travel with his wife Laura to observe cuckoos and record their songs in the field. Karen Klitz began painting birds in Africa in 1965. She has been a staff illustrator in the Museums of Zoology and Paleontology at the University of Michigan and is currently Principal Illustrator for the Museum of Vertebrate Zoology at the University of California, Berkeley.

Bird Families of the World 1. The Hornbills Alan Kemp 2. The Penguins Tony D.Williams 3. The Megapodes Darryl N. Jones, René W. R. J. Dekker, and Cees S. Roselaar 4. Fairy-wrens and Grasswrens Ian Rowley and Eleanor Russell 5. The Auks Anthony J. Gaston and Ian L. Jones 6. The Birds of Paradise Clifford B. Frith and Bruce Beehler 7. The Nightjars and their Allies D.T. Holyoak 8. Toucans, Barbets and Honeyguides Lester L. Short and Jennifer F. M. Horne 9. Ratites and Tinamous S. J. J. F. Davies 10. The Bowerbirds Clifford B. Frith and Dawn W. Frith 11. Albatrosses and Petrels across the World Michael Brooke 12. The Grebes Jon Fjeldså 13. The Hawaiian Honeycreepers H. Douglas Pratt 14. Herons James A. Kushlan and James A. Hancock 15. The Cuckoos Robert B. Payne 16. Ducks, Geese and Swans Edited by Janet Kear

Bird Families of the World

The Cuckoos Robert B. Payne with a molecular genetic analysis of cuckoo phylogeny by

Michael D. Sorenson and Robert B. Payne Color plates by

Karen Klitz Black and white illustrations by

John Megahan

Great Clarendon Street, Oxford OX2 6DP Oxford University Press is a department of the University of Oxford. It furthers the University’s objective of excellence in research, scholarship, and education by publishing worldwide in Oxford New York Auckland Cape Town Dar es Salaam Hong Kong Karachi Kuala Lumpur Madrid Melbourne Mexico City Nairobi New Delhi Taipei Toronto Shanghai With offices in Argentina Austria Brazil Chile Czech Republic France Greece Guatemala Hungary Italy Japan South Korea Poland Portugal Singapore Switzerland Thailand Turkey Ukraine Vietnam Oxford is a registered trade mark of Oxford University Press in the UK and in certain other countries Published in the United States by Oxford University Press Inc., New York Illustrations © Oxford University Press, 2005 Text © Robert B. Payne except Text of Molecular Systematics Chapter © Michael D. Sorenson and Robert D. Payne, 2005 The moral rights of the authors have been asserted Database right Oxford University Press (maker) First published 2005 All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, without the prior permission in writing of Oxford University Press, or as expressly permitted by law, or under terms agreed with the appropriate reprographics rights organization. Enquiries concerning reproduction outside the scope of the above should be sent to the Rights Department, Oxford University Press, at the address above You must not circulate this book in any other binding or cover and you must impose this same condition on any acquirer A catalogue record for this title is available from the British Library British Library Cataloging in Publication Data Data available ISBN 0 19 850213 3 10 9 8 7 6 5 4 3 2 1 Typeset by Macmillan India Limited Printed in China

Acknowledgments

My father Russell Payne and Sao Saimong Mongrai introduced me to cuckoos in the Shan States in Burma in 1951, where I watched cuckoos at Taunggyi, a site of classic cuckoo studies in the 1930s. Undergraduate work at the University of Michigan Museum of Zoology and a talk by Herbert Friedmann on brood-parasitic birds at Cranbrook Institute of Science were important in turning attention toward the cuckoos. Graduate work at the Museum of Vertebrate Zoology, University of California, Berkeley, supported a field study of broodparasitic cowbirds. A National Science Foundation postdoctoral research fellowship in 1965 allowed two years in the field with brood-parasitic cuckoos in Africa, with the help of Herbert Friedmann and the Percy FitzPatrick Institute of African Ornithology at the University of Cape Town, with fieldwork in South Africa, Botswana, Rhodesia, Zambia, Malawi and Kenya. In southern and east Africa, Karen Klitz and I worked with cuckoos in the 1960s, guided by Jack Winterbottom, Richard Liversidge, Jack Skead and Gordon Ranger in South Africa, Michael Irwin in Zimbabwe, and John Williams in Kenya. My wife Laura Payne and I watched Neotropical cuckoos in Gayana and Curaçao, we next shared fieldwork with cuckoos and fairy-wrens in the 1980s in Western Australia with the CSIRO research team of Ian Rowley, Eleanor Russell, Michael Brooker and Graeme Chapman on Gooseberry Hill, and with Ian and Eleanor and Dick and Molly Brown in the Manjimup forests. Later we worked with Nedra Klein in Cameroon, Mike Sorenson in Zimbabwe and Guinea, Pierre Reynaud in Senegal, Clive Barlow in Guinea, Senegal, Gambia and Mali, Ben King in India, Phil Round in Thailand,Wang Luan Keng

and Lim Kim Keang in Malaysia, Ir Darjono in Java, and Kathy Groschupf in Arizona. In other field studies I worked with Geoffrey Field in Sierra Leone, Chris Risley in Ghana, Kathy Groschupf in Cameroon, and Roger Wilkinson, Neville Skinner and Mark Hopkins in Nigeria. In North America the University of Michigan Biological Station, the Matthai Botanical Gardens, and the E. S. George Reserve of the University of Michigan provided field support, as did the Oklahoma University Biological Station at Willis, Oklahoma. The second phase of work was examining the specimens in natural history museums.This gave a closer look at cuckoos than the birds themselves ever allowed in the field. The curators of these valuable collections allowed me to examine specimens during research visits or through loans. For permission to examine specimens and information from their collections, I thank the curators and staff of the following museums and live bird collections: Brian Gill (AM,Auckland); Dean Amadon, George Barrowclough, Joel Cracraft, Wesley Lanyon, Mary LeCroy, Pamela Beresford, Peter Capainolo, Shannon Kenney, Chris Milensky, Chris Vogel and Paul Sweet (AMNH, New York); David Agro and Leo Joseph (ANSP, Philadelphia); Robert PrysJones, Michael Walters, Frank Steinbrenner and Mark Adams (BMNH,Tring); Carla Kishinami (B. P. Bishop Museum, Honolulu); Kit Hustler and Michael Irwin (BWYO, Bulawayo); Phil Bruner (BYU, Laie, Hawaii); Lei Fu-min (CAS, Beijing); Ken Parkes, Brad Livezey and Robin Panza (CM, Pittsburgh); Robert S. Kennedy (CMNH, Cincinnati); Michael Brooke (Cambridge University Museum); Kevin McGowan and Charles Dardia

vi Acknowledgments (CU, Ithaca); Mike Brooke (CUM, Cambridge); David Allan and Philip Clancey (DM, Durban); Gene Hess and Jean Woods (DMNH, Wilmington); Dave Willard, John Bates, Shannon Hackett, Melvin Traylor, Peter Lowther and Amy Driskell (FMNH, Chicago); Dave Steadman and Tom Webber (FLMNH, Gainesville); René Verheyen (IRSNB, Brussels), Bob Zink, Scott Lanyon and John Klicka ( JFBM, Minneapolis); C. J. Skead (Kaffrarian Museum, King William’s Town); H. Schouteden, Anton De Roo and Michel Louette (KMMA, Tervuren); Rick Prum and Mark Robbins (KU, Lawrence); Donna Dittmann, Steve Cardiff, Rob Moyle, Van Remsen and Fred Sheldon (LSU, Baton Rouge); Karl Schuchmann, Renate van den Elzen and Hans Wolters (ZFMK, Bonn); Eulàlia Garcia Franquesa (MCNB, Barcelona); Doug Causey, Ray Paynter and Alison Pirie (MCZ, Cambridge); Robin Restall (MHNLS); Pavel Tomkovich (MMZ, Moscow); Christian Erard and C. Voisin (MNHN, Paris); Martin Nhlane (MNM, Blantyre); Giuliano Doria, Roberto Poggi and Enrico Borgo (MSNG, Genoa); Giorgio Chiozzi and Carlo Violani (MSNM, Milan); Laura Abraczinskas, Barb Lundrigan and Pamela Rasmussen (MSU, East Lansing); Ned Johnson, Carla Cicero and Karen Klitz (MVZ, Berkeley); S. Somadikarta and Ir Darjono (MZB Cibinong); Marta Poggesi (MZUS, Firenze); Tony Parker (NMGL, Liverpool); Chum Cunningham-van Someren (NMK, Nairobi); Rodolfo A. Caberoy (NMP, Manila); Ernst Baurenfeind and Herbert Schifter (NMW, Wien); Jarunjin Nabhitabhata (NSM, Bangkok); Gary Schnell (OU); Robin Restall (POC); Steve van Dyke (QM, Brisbane); René Dekker, Gerlof Mees and Peter Van Dam (RMNH, Leiden); Jon Barlow and Jim Dick (ROM, Toronto); John M. Winterbottom (South African Museum, Cape Town); Gerald Mayr (SMF, Frankfurt); Siegfried Eck (SMTD, Dresden), Maklarin Bin Lakim (SP, Kinabalu); James Mueller (SRSU); Keith Arnold (TCWC, College Station, Texas); Lakkana Pakkarnseree (TISTR, Bangkok); Fritz Hertel (UCLA, Los Angeles); Bob Dickerman (UNM, Albuquerque); Gary Voelcker (UNLV, Las Vegas); Mike Braun, Phil Angle, James Dean, Gary Graves, Helen James, Storrs Olson, Pamela Rasmussen and

Dick Zusi (USNM,Washington, D.C.); Scott Cutler (UTEP); Scott Edwards, Sievert Rohwer,Wang Luan Keng, Chris Wood, Gary Voelcker, Robert Faucett, Sharon Birks and Sergei Drovetski (UWBM, Seattle); Dieter Rinke (VW, Walsrode); Glen Storr (WAM, Perth); René Corrado (WFVZ, Camarillo); Jacques Gautier, Paul Whitfield and Kristof Zyskowski (YPM, New Haven); Juergen Fiebig (ZMB, Berlin); Jon Fjeldså (ZMUC, Copenhagen); Yang Chang Man, Darren C. J. Yeo and Lua Hui Keng (ZRC, Singapore); Josef Reichholf (ZSM, München); Kees Roselaar (ZMA,Amsterdam); and Pavel Tomkovich (ZMM, Moscow). Tom Huels hand-reared young roadrunners for developmental studies. Michael Walters measured eggs at BMNH. Dan Gaebel, Karen Kevelighan, Jeff Rebitzke, Sharon Reske and Angie Steffen, undergraduate students at the University of Michigan, examined skeletal specimens with me. George Kulesza made available the measurements of brain size in cuckoos from his comparative brain measurements from skeletons in UMMZ. Carlo Violani was helpful in arranging visits to natural history museums in Italy and in locating specimens and publications. Godfrey Symons gave access to photograph bird eggs in his collection; the collection is now in the Cathcart Museum, South Africa, and Jean Woods at DMNH loaned bird eggs to use in calibrating egg colors in the photographs. For photographs of museum specimens, I thank Patricia Escalante Pliego and Adolfo G. Navarro S. (Colección Nacional de Aves, Dept. Zoología, Instituto de Biología, UNAM, Mexico City), Leslie Jessop (NEWHM), Michel Louette (KMMA), Maklarin Bin Lakim (Sabah Parks), Ir Darjono (MZB) and Lei Fu-min (CAS), and for photographs of live cuckoos I thank Graeme Chapman, Carla Fontana, Johannes Foufopoulos, Karen Klitz, Mary LeCroy, Dale Lewis, Regina Macedo, Richard Peek, Cheong Weng Chuan and Kit Hustler, Ron Orenstein, Ole Post, Ong Kiem Sian, Trevor Price, Carol Ralph, Cheong Wan Chun, Pamela Rasmussen, Phil Round, Joe Strauch, Bob Thornton, Sandra Vehrencamp, Wang Luan Keng, Jongmin Yoon and Moo-Boo Yoon. Rob Moyle and Carl Vernon provided their photographs of forest habitat which we adapted into the background of the color plates.The British Library National Sound Archive (NSA, Richard

Acknowledgments vii Ranft), Cornell Library of Natural Sounds (LNS), Sandra Gaunt (OSU), Per Alström, Clive Barlow, Graeme Chapman, Jared Diamond, Chris Filardi, Lei Fumin, Paul Holt, Bob Kennedy, Ben King, Lindy McBride, Barry Morgan, Pamela Rasmussen, Phil Round, Pratap Singh, Deepal Warakagoda and Martha Whitson gave descriptions of songs and copies of their tape and digital recordings of cuckoo songs. Colleagues at the University of Michigan, including Lei Fumin, Janet Hinshaw, George Kulesza, David Mindell, John Mitani, Phil Myers, Robert Storer, Jean Woods and Tamaki Yuri, were helpful throughout the study. Material for molecular genetics analysis was provided by several museums and collections (AMNH, ANSP, BMNH, CM, CMNH, FMNH, LSU, QM, AM, DM, DMNH, MCZ, MNM, MVZ, RMNH, SMNS, UMMZ, UMMZ, USNM, UWBM, VW, WFVZ, YPM and ZMUC), and by Dinesh Bhatt (Gurukul Kangri University, Haridwar), Steve Latta (University of Missouri), Terry Root (Ann Arbor), Steve Schneider (Palo Alto) and Lei Fu-min (CAS, Beijing). Preliminary genetic analyses were conducted in David Mindell’s laboratory at the University of Michigan Museum of Zoology, for which the University of Michigan supported funding for DNA sequencing. Suzanne Ambs, Marina Ramon, Tracy Heath, Kristina Sefc and Elen Oneal assisted with the processing of DNA sequencing reactions at Boston University. Funding for DNA sequencing and a computer cluster used for the phylogenetic analyses were provided by Boston University. For bibliographic help I thank Joann Constantinides, Janet Hinshaw, Charlene Stachnick and Dorothy Riemenschneider (University of Michigan, Ann Arbor), Linda Birch (Alexander Library, EGI, Oxford University), Effie Wahr (BMNH, Tring), Mary Petersen (ZMUC, Copenhagen), Renate van den Elzen (ZFMK, Bonn) and the librarians of ZSM (Munich). David Seel of Bangor guided me to Pound Green Common, Worcestershire, where Edgar Chance observed cuckoos in the 1920s. Carla Fontana, Lydia Skrynnikova and Tamaki Yuri translated works in Portuguese, Russian and Japanese; and Ir Darjono, James Dean, Antonia Gorog and Paul Taylor interpreted the museum specimen labels of birds taken in Indonesia. David Allan, Wayne Arendt, Walter Boles, Mariana

Cariello, Mario Cohn-Haft, Nigel Collar, Jack Cox, Carla Fontana, Larry Heaney, Regina Macedo, Sergio Posso, Pamela Rasmussen, Danny Rogers, Burkhard Stephan and Robert L. Sutton kindly sent original data, observations and copies of their papers, and Burkhard Stephan allowed us to use his drawings of cuckoo wings. Clive Barlow, Pamela Beresford, Andy Berger, Bonnie Bowen, Michael Brooker, John Colebrook-Robjent, Richard Dean, Edward Dickinson, Jon Fjeldså, Steve Goodman, Kathy Groschupf, Bill Hamilton, Hiroyoshi Higuchi, Tom Huels, Kit Hustler, Klaus Immelmann, Bob Kennedy, Jiro Kikkawa, J. M. (Martjan) Lammetink, Mary LeCroy, Regina Macedo, Gerald Mauersberger, Ernst Mayr, Sergio Posso, Pamela Rasmussen, Manolo Soler, Richard Thorington, Carl Vernon, Carlo Violani, Michael Walters and Amotz Zahavi offered discussion during the study. For comments on the manuscript I thank Bonnie Bowen, Richard Dean, Edward Dickinson, Kathy Groschupf, Ben King, Karen Klitz, George Kulesza, David Lahti, Regina Macedo, Laura Payne, Rick Prum, Pamela Rasmussen, Robin Restall, Charlie Rosa, Mike Sorenson and J.Van Remsen. Karen Klitz painted the color plates of cuckoos. John Megahan drew the black and white illustrations of birds and bones and digitized the maps, the eggs and nestling photographs. Janet Hinshaw, Alec Lindsay, David Lahti, John Megahan, Laura Payne, Jean Woods and Tamaki Yuri helped prepare the illustrations. The National Science Foundation supported fieldwork on cuckoos and laboratory work on molecular genetics. Additional support was provided for field studies by the National Geographic Research Committee and the University of Michigan, and for molecular studies by the University of Michigan Museum of Zoology and Boston University. I thanks the series editors Professors Christopher Perrins,Walter Bock, and Jiro Kikkawa for their support, Judith May, formerly Senior Editor, for inviting the contribution, and the staff of Oxford University Press and Macmillan India, especially Stephen Benaim and Laura O’Neill, all for providing the freedom and encouragement to develop the book. I thank OUP for publishing this wonderful series on birds.

A brood parasite destroys its host and family—speaks the Fool, as Goneril evicts Lear: “For you know, nuncle, The hedge-sparrow fed the Cuckoo so long That it’s had its head bit off by its young. (King Lear, I, 4, 203–205),

“. . . the Cuckoos. In these the confusion is tremendous, and I don’t know if [I] will be able to make them out satisfactorily. I am afraid that I shall have not the time to go through them in a proper way.”— Tommaso Salvadori, 1875, letter to R. B. Sharpe, 12 Oct 1877 (Violani et al. 1997: 40–41).

Laying behavior of Little Bronze-cuckoo Chrysococcyx minutillus in Putrajaya Wetland, near Kuala Lumpur, Malay Peninsula. a, female cuckoo perches and peers at a pocket-shaped nest of a flyeater, Golden-bellied Gerygone Gerygone sulphurea. The cuckoo then perched closer to the nest, then moved to the nest. b, cuckoo laying in nest, perched on nest pocket rim, the head and cloacal region inserted into the nest entrance, the back feathers raised, the wings and tail spread outside the nest, wings grasping the nest and the tail supporting the bird (obscured by vegetation). The cuckoo remained in this posture for 3–4 seconds. After the cuckoo's visit, an unspotted greenish egg appeared in the nest — Little Bronze-cuckoos lay greenish eggs; the flyeaters lay spotted whitish eggs. Drawn from photos by Cheong Weng Chun. The laying behavior of this cuckoo is similar to that described in other bronze-cuckoo species in Australia (Brooker et al. 1988).

Contents

1 2 3 4 5 6 7 8 9 10 11

List of color plates List of abbreviations Plan of the book

xiv xv xix

PART I General chapters Introduction to the cuckoos Distribution, habitats and conservation status Behavior Morphology A molecular genetic analysis of cuckoo phylogeny Species Fossil and comparative evidence of cuckoo relationships Breeding biology and life histories Cooperative breeding Brood parasitism The evolution of brood parasitism in cuckoos

3 9 15 27 68 95 109 114 132 137 154

PART II Species accounts Crotophaginae Genus Guira Guira Cuckoo Genus Crotophaga Greater Ani Smooth-billed Ani Groove-billed Ani

169 169 169 172 172 174 178

Neomorphinae Genus Tapera American Striped Cuckoo Genus Dromococcyx Pheasant Cuckoo Pavonine Cuckoo Genus Morococcyx Lesser Ground-cuckoo Genus Geococcyx Greater Roadrunner Lesser Roadrunner

Guira guira Crotophaga major Crotophaga ani Crotophaga sulcirostris

Tapera naevia Dromococcyx phasianellus Dromococcyx pavoninus Morococcyx erythropygus Geococcyx californianus Geococcyx velox

183 183 183 187 187 189 191 191 193 193 198

x Contents Genus Neomorphus Banded Ground-cuckoo Rufous-winged Ground-cuckoo Red-billed Ground-cuckoo Rufous-vented Ground-cuckoo Centropodinae Genus Centropus Buff-headed Coucal Pied Coucal Greater Black Coucal Biak Coucal Rufous Coucal Green-billed Coucal Black-faced Coucal Black-hooded Coucal Short-toed Coucal Bay Coucal Gabon Coucal Black-throated Coucal Senegal Coucal Blue-headed Coucal Coppery-tailed Coucal White-browed Coucal Javan Coucal Greater Coucal Goliath Coucal Madagascar Coucal African Black Coucal Philippine Coucal Lesser Coucal Violaceous Coucal Lesser Black Coucal Pheasant Coucal Couinae Genus Carpococcyx Sumatran Ground-cuckoo Bornean Ground-cuckoo Coral-billed Ground-cuckoo Genus Coua Crested Coua Verreaux’s Coua Blue Coua Red-capped Coua Red-fronted Coua Coquerel’s Coua

radiolosus rufipennis pucheranii geoffroyi

200 200 202 203 204

milo ateralbus menbeki chalybeus unirufus chlororhynchos melanops steerii rectunguis celebensis anselli leucogaster senegalensis monachus cupreicaudus superciliosus nigrorufus sinensis goliath toulou grillii viridis bengalensis violaceus bernsteini phasianinus

208 208 208 210 211 213 214 215 216 218 219 221 222 224 226 228 231 233 236 238 242 244 246 248 250 254 255 257

Neomorphus Neomorphus Neomorphus Neomorphus

Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus Centropus

Carpococcyx viridis Carpococcyx radiatus Carpococcyx renauldi Coua Coua Coua Coua Coua Coua

cristata verreauxi caerulea ruficeps reynaudii coquereli

262 262 262 264 266 268 268 270 271 273 274 276

Contents xi Running Coua Giant Coua Snail-eating Coua Red-breasted Coua

Cuculinae Genus Rhinortha Raffles’s Malkoha Genus Ceuthmochares Whistling Yellowbill Chattering Yellowbill Genus Taccocua Sirkeer Malkoha Genus Zanclostomus Red-billed Malkoha Genus Phaenicophaeus Chestnut-breasted Malkoha Chestnut-bellied Malkoha Red-faced Malkoha Blue-faced Malkoha Black-bellied Malkoha Green-billed Malkoha Genus Rhamphococcyx Yellow-billed Malkoha Genus Dasylophus Rough-crested Cuckoo Scale-feathered Malkoha Genus Clamator Chestnut-winged Cuckoo Great Spotted Cuckoo Levaillant’s Cuckoo Jacobin Cuckoo Genus Coccycua Little Cuckoo Dwarf Cuckoo Ash-colored Cuckoo Genus Piaya Squirrel Cuckoo Black-bellied Cuckoo Genus Coccyzus Dark-billed Cuckoo Yellow-billed Cuckoo Pearly-breasted Cuckoo Mangrove Cuckoo Cocos Cuckoo Black-billed Cuckoo Gray-capped Cuckoo

Coua Coua Coua Coua

cursor gigas delalandei serriana

277 279 280 281

283 283 Rhinortha chlorophaea 283 285 Ceuthmochares australis 285 Ceuthmochares aereus 287 291 Taccocua leschenaultii 291 293 Zanclostomus javanicus 293 294 Phaenicophaeus curvirostris 294 Phaenicophaeus sumatranus 298 Phaenicophaeus pyrrhocephalus 299 Phaenicophaeus viridirostris 301 Phaenicophaeus diardi 302 Phaenicophaeus tristis 304 306 Rhamphococcyx calyorhynchus 306 308 Dasylophus superciliosus 308 Dasylophus cumingi 309 311 Clamator coromandus 311 Clamator glandarius 313 Clamator levaillantii 318 Clamator jacobinus 320 325 Coccycua minuta 325 Coccycua pumila 328 Coccycua cinerea 330 331 Piaya cayana 331 Piaya melanogaster 335 336 Coccyzus melacoryphus 336 Coccyzus americanus 339 Coccyzus euleri 345 Coccyzus minor 347 Coccyzus ferrugineus 350 Coccyzus erythropthalmus 351 Coccyzus lansbergi 354

xii Contents Chestnut-bellied Cuckoo Rufous-breasted Cuckoo Jamaican Lizard-cuckoo Puerto Rican Lizard-cuckoo Cuban Lizard-cuckoo Hispaniolan Lizard-cuckoo Genus Pachycoccyx Thick-billed Cuckoo Genus Microdynamis Dwarf Koel Genus Eudynamys Common Koel Genus Urodynamis Long-tailed Cuckoo Genus Scythrops Channel-billed Cuckoo Genus Chrysococcyx Asian Emerald Cuckoo Violet Cuckoo Diederik Cuckoo Klaas’s Cuckoo Yellow-throated Cuckoo African Emerald Cuckoo Long-billed Cuckoo Horsfield’s Bronze-cuckoo Black-eared Cuckoo Rufous-throated Bronze-cuckoo Shining Bronze-cuckoo White-eared Bronze-cuckoo Little Bronze-cuckoo Genus Cacomantis Pallid Cuckoo White-crowned Cuckoo Chestnut-breasted Cuckoo Fan-tailed Cuckoo Banded Bay Cuckoo Plaintive Cuckoo Gray-bellied Cuckoo Brush Cuckoo Genus Cercococcyx Dusky Long-tailed Cuckoo Olive Long-tailed Cuckoo Barred Long-tailed Cuckoo Genus Surniculus Fork-tailed Drongo-cuckoo Philippine Drongo-cuckoo Square-tailed Drongo-cuckoo

Coccyzus Coccyzus Coccyzus Coccyzus Coccyzus Coccyzus

pluvialis rufigularis vetula vieilloti merlini longirostris

Pachycoccyx audeberti Microdynamis parva Eudynamys scolopacea Urodynamis taitensis Scythrops novaehollandiae Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Chrysococcyx Cacomantis Cacomantis Cacomantis Cacomantis Cacomantis Cacomantis Cacomantis Cacomantis

maculatus xanthorhynchus caprius klaas flavigularis cupreus megarhynchus basalis osculans ruficollis lucidus meyeri minutillus

pallidus leucolophus castaneiventris flabelliformis sonneratii merulinus passerinus variolosus

Cercococcyx mechowi Cercococcyx olivinus Cercococcyx montanus Surniculus dicruroides Surniculus velutinus Surniculus lugubris

356 357 358 360 361 362 364 364 367 367 368 369 378 378 380 381 385 385 388 391 395 399 400 403 405 407 409 410 413 415 421 422 424 426 427 430 433 437 439 447 447 449 451 453 454 458 460

Contents xiii Moluccan Drongo-cuckoo Genus Hierococcyx Mustached Hawk-cuckoo Dark Hawk-cuckoo Large Hawk-cuckoo Common Hawk-cuckoo Rufous Hawk-cuckoo Philippine Hawk-cuckoo Javan Hawk-cuckoo Whistling Hawk-cuckoo Genus Cuculus Black Cuckoo Red-chested Cuckoo Asian Lesser Cuckoo Sulawesi Cuckoo Indian Cuckoo Madagascar Lesser Cuckoo African Cuckoo Oriental Cuckoo Himalayan Cuckoo Sunda Lesser Cuckoo Common Cuckoo

Surniculus musschenbroeki Hierococcyx Hierococcyx Hierococcyx Hierococcyx Hierococcyx Hierococcyx Hierococcyx Hierococcyx Cuculus Cuculus Cuculus Cuculus Cuculus Cuculus Cuculus Cuculus Cuculus Cuculus Cuculus

vagans bocki sparverioides varius hyperythrus pectoralis fugax nisicolor

clamosus solitarius poliocephalus crassirostris micropterus rochii gularis optatus saturatus lepidus canorus

463 464 465 466 468 471 473 475 477 479 481 481 485 488 491 492 496 498 500 505 508 510

Glossary

519

Bibliography

526

Index

607

Color plates Color plates fall between pages 166–167. Plate 1 Plate 2 Plate 3 Plate 4 Plate 5 Plate 6 Plate 7 Plate 8 Plate 9 Plate 10 Plate Plate Plate Plate Plate Plate Plate

11 12 13 14 15 16 17

Plate 18 Plate 19 Plate 20

New World cuckoos Crotophaga, Guira,Tapera, Dromococcyx and Morococcyx New World Geococcyx and Neomorphus Coucals Centropus milo, ateralbus, menbeki, chalybeus, unirufus, celebensis, melanops, steerii and rectunguis Coucals Centropus nigrorufus, sinensis, chlororhynchos, phasianinus (including nigricans and spilopterus) and bernsteini Coucals Centropus toulou, violaceus, goliath, bengalensis, viridis, grillii African coucals Centropus leucogaster, anselli, monachus, cupreicaudus, senegalensis and superciliosus Old World ground-cuckoos Carpococcyx and couas Coua Old World malkohas Ceuthmochares, Rhinortha, Taccocua, Zanclostomus, Rhamphococcyx, Dasylophus and Phaenicophaeus New World cuckoos Coccycua, Piaya and Coccyzus (including Saurothera and Hyetornis) Old World brood-parasitic genera Pachycoccyx, Microdynamis, Eudynamys, Urodynamis and Scythrops Glossy cuckoos in Africa and Asia, genus Chrysococcyx Glossy cuckoos in Australasia, genus Chrysococcyx Brush cuckoos Cacomantis Drongo-cuckoos Surniculus, long-tailed cuckoos Cercococcyx Crested cuckoos Clamator, hawk-cuckoos Hierococcyx Cuckoos Cuculus Eggs of brood-parasitic cuckoos and their hosts in Australia and southern Africa. Brood parasitism in cuckoos. Eggs of Village Weaver Ploceus cucullatus Mouth patterns of young cuckoos

Abbreviations AM AMNH ANSP BMNH BPBM BWYO BYU CAS CM CMNH CU CUM DM DMNH FLMNH FMNH IRSNB JFBM KMMA KU LNS LSU MCNB MCZ MHNLS MMZ MNHN MNM MSNG MSNM MSU MVZ MZB MZUS NMGL NMK NMP NMW NSA

Auckland Museum, Auckland, New Zealand American Museum of Natural History, New York Academy of Natural Sciences, Philadelphia British Museum of Natural History (Natural History Museum),Tring, UK Bernice P. Bishop Museum, Honolulu National Museums of Zimbabwe, Bulawayo Museum of Natural History, Brigham Young University, Hawaii Campus, Laie, Hawaii Chinese Academy of Sciences, Beijing Carnegie Museum, Pittsburgh Cincinnati Museum of Natural History, Cincinnati Cornell University, Ithaca Cambridge University Museum, Cambridge Durban Museum, Durban Delaware Museum of Natural History, Greenville Florida State Museum of Natural History, Gainesville Field Museum of Natural History, Chicago Institut Royal des Sciences Naturales de Belgique, Brussels James Ford Bell Museum of Natural History, University of Minnesota Koninklijk Museum voor Midden Afrika,Tervuren University of Kansas Museum of Natural History, Lawrence Library of Natural Sounds, Cornell University, Ithaca, New York Museum of Zoology, Louisiana State University, Baton Rouge Museu de Cièncias Naturals de la Ciutadella, Barcelona Museum of Comparative Zoology, Cambridge, MA Sociedad de Ciencias Naturales La Salle (Fondacion La Salle), Caracas Moscow Museum of Zoology, Moscow Muséum National d’Histoire Naturelle, Paris Malawi National Museum, Blantyre Museo Civico di Storia Naturale “Giacomo Doria”, Genoa Museo Civico di Storia Naturale, Società Italiana di Scienze Naturale, Milan Michigan State University Museum, East Lansing Museum of Vertebrate Zoology, University of California, Berkeley Museum Zoologicum Bogoriense, Cibinong nr Bogor Museo Zoologico Universitario “La Specola”, Florence National Museums and Galleries (Merseyside Museums), Liverpool National Museum of Kenya, Nairobi National Museum of the Philippines, Manila Naturhistorisches Museum Wien,Vienna British Library National Sound Archive, London

xvi Abbreviations NSM OU OSU POC QM RMNH ROM SMF SMNS SMTD SP SRSU TCWC TISTR UCLA UMMZ UNLV UNM USNM UTEP UWBM WAM WFVZ YIO YPM VW ZFMK ZMA ZMB ZMM ZMUC ZRC ZSM F I, Is M NP NSW NT P PN S

National Science Museum,Technopolis, Bangkok University of Oklahoma, Norman Borror Laboratory of Bioacoustics, Ohio State University, Columbus Phelps Ornithological Collection, Caracas,Venezuela Queensland Museum, Brisbane Nationaal Natuurhistorisch Museum, Leiden Royal Ontario Museum,Toronto Forschungsinstitut Senckenberg, Frankfurt Staatliches Museum für Naturkunde in Stuttgart, Stuttgart Staatliches Museum für Tierkunde, Dresden Sabah Parks, Kota Kinabalu Scudday Vertebrate Collection, Sul Ross State University, Alpine,Texas Texas Cooperative Wildlife Collection, College Station Thailand Institute of Scientific and Technical Research, Bangkok Dickey Collection, University of California, Los Angeles University of Michigan Museum of Zoology, Ann Arbor University of Nevada at Las Vegas, Las Vegas (MBM) University of New Mexico Museum of Southwestern Biology, Albuquerque National Museum of Natural History,Washington, D.C. University of Texas at El Paso, El Paso Burke Museum, University of Washington, Seattle Western Australian Museum, Perth Western Foundation of Vertebrate Zoology, Camarillo, California Yamashina Institute of Ornithology, Abiko, Japan Yale Peabody Museum, New Haven Vogelpark Walsrode,Walsrode, Germany Museum Alexander Koenig, Bonn Universiteit van Amsterdam Museum, Amsterdam Zoologische Museum, Humboldt-Universitat, Berlin Zoological Museum of Moscow, Moscow Universitets Zoologiske Museum, Copenhagen National Museum of Singapore (Zoological Reference Collection, National University of Singapore), Singapore Zoologische Staatssammlung München, Munich female island(s) male National Park New South Wales, Australia Northern Territory, Australia wing feather, primary (numbered from innermost outward, P1 the inner primary, P10 the outer primary) Parc National wing feather, secondary (numbered from outermost inward)

Abbreviations xvii T TA TB U WA

tail feather, rectrix (numbered from the inner pair outward,T1 the inner rectrix,T5 the outer rectrix) ambient temperature, air temperature body temperature unsexed Western Australia

This page intentionally left blank

Plan of the book

The book is a comprehensive survey of the biology of cuckoos, with descriptions of the life history and systematic status of the 141 cuckoo species. Part I gives general accounts of cuckoo biology, behavior, cooperative breeding and brood parasitism, and it develops ideas about the relationships among the species and major lineages of cuckoos, based on song, morphology and molecular genetics. In the general chapters the cuckoos are mentioned by common name or scientific name, and sometimes by both names. In part I the text, tables and graphs summarize the details of life history and comparative morphology of the cuckoos. Part II consists of species accounts of each cuckoo species, with details about plumage, size, geographic variation, behavior and breeding biology, depending on the information available.The details and references for the general chapters in Part I are in the species accounts in Part II. Recent field observations on the behavior of these birds and phylogenetic studies on cuckoo evolution, allow a synthesis of ideas about the evolution of parental behavior and adaptations for competition, cooperation and brood parasitism in these remarkable birds. The species were determined by the morphological distinctiveness of regional populations and by their songs when these were available on tape recordings. Each cuckoo species that has been studied and is well known has a characteristic song.The songs provide an objective criterion to assess the status of species: if cuckoos have the same song they are conspecific, and if the songs differ, the cuckoos are not of the same species. Female birds choose their mates on the basis of song. Females are attracted to songs of their own species and not to songs of other

species. For these reasons bird songs are a major clue to the nature and limits of biological species (Payne 1986, Alström and Ranft 2003). Another criterion used to determine species status is the fact that birds living in the same area and not interbreeding are distinct species. So song is used mainly to compare birds living in different areas, such as in continental Asia and on islands. In a few cases the results of molecular genetics were used to refine ideas about the nature and limits of cuckoo species. The scientific and common names of most birds in the text are the names in Sibley and Monroe (1990), a comprehensive recent and referenced systematic revision of birds on a worldwide scale.Where the names in that source differ from those in standard regional works, as in Japan, the regional names are also included. For cuckoos the names in Sibley and Monroe’s book are used, except where recent evidence of song or molecular genetics, or both, indicates that the species are not the ones those authors suggested. For example, in the hawk-cuckoos Hierococcyx fugax and the geographic forms nisicolor, hyperythrus and pectoralis, there now appear to be four species rather than one ‘Hodgson’s Hawkcuckoo’. In Part 1 the common and scientific names are given together at their first occurrence in the text. Subsequently the common names are used, except in the tables and chapters that discuss systematic details, where the scientific names are used again. For most cuckoo species and subspecies, 20 or more specimens were measured if the series were available in UMMZ, otherwise specimens in other museums were sampled. When possible, the sex of the specimens used was determined when the

xx Plan of the book gonads were examined, and male or female indicated on the label. In a few specimen series it became obvious that the birds had been mis-sexed. Some species are now known to differ in male and female plumage, which was not known when the collector obtained the birds, for instance with Raffles’s Malkoha Rhinortha chlorophaea (Swainson 1838, Shelley 1891). In other species such as coucals Centropus the sexes differ in size. I avoided museum series where mis-sexing of birds was apparent. In some cases the sex had been added after the field label was written. In such cases I usually did not use the specimen, although some may have been labeled appropriately from the collector’s field notes after the field label was written. The publications where each species was first described are listed in the species accounts. Citations of publications where subspecies were described are available elsewhere (Peters 1940, Payne 1997b), and for recently described forms the source is indicated in the species account. Measurements were taken on museum specimens. Wing length was measured with a stop-end mm rule and the wing flattened on the rule. Tail length was taken with a rule from the point of insertion of the central rectrices (tail feathers) on the body to the tip of the central tail feather. The tail could not be measured on some specimens of which the collector had pushed the base of the tail well into the body cavity, as far as 100 mm in the large coucals and malkohas that were collected and prepared by G. Heinrich, as described by Eck (1978). The tail feathers are numbered in the text, T1 being the innermost rectrix,T2 the next, and T5 the outermost rectrix. Bill length was measured with calipers from the skull to the tip of the bill. Tarsus length was measured from the proximal end of the tarsometatarsus to the distal margin of the most distal tarsal scute that was not divided between toes 2, 3 and 4. Specimens were measured for all species and nearly all subspecies known to vary in size, and in others that vary in plumage color or pattern. For cuckoos that I did not measure, I used the measurements given by E. Mayr, E. Stresemann, D. Rogers in Higgins (1999), where these coincided with the measurements of specimens, done by other ornithologists. Bird skins

shrink as they dry, and the wing length of museum specimens is less than that of live birds measured in the field, with a difference of about 2 mm in the Yellow-billed Cuckoo Coccyzus americanus. Measurements are given in terms of maximum, minimum, mean, and ⫾ standard deviation (sd) where sample sizes are larger (n ⫽ 6 or more).Where data are cited from another author, sd is included only when the author included this index of variation. Body weight is the main index of body size in birds. Body weight can change as much as 40% in resident cuckoos from summer to winter (Greater Roadrunner Geococcyx californianus), more than 20% above average for breeding females with a large yolking ovary and oviduct, 70% above average for birds that had large stores of fat before migration, and 40% below average after a long over-water migration (Yellow-billed Cuckoos Coccyzus americanus). Individual birds that are caught and weighed through the year even in a tropical region change their body weight by as much as 16% (Mustached Hawk-cuckoo Hierococcyx vagans).Where specimen labels noted a large amount of premigratory fat, or laying females with enlarged ovaries and oviducts and sometimes an egg in the oviduct, the body weights of these specimens were noted separately. Body weights do not vary in a statistically normal frequency distribution and for this reason sd’s are not included for weights. The wing formula captures some of the diversity in wing shape. Counting the primaries from the innermost outward, the rank of feather length from the bend of the wing to the tip of the folded wing was determined, with the longest primary listed first, followed by the next, and so on. For example, P7 ⬎ 6 ⬎5 indicates that the seventh primary (P7) is longest, then P6 and then P5. Cuckoos with the longest primary number 8 and 9 have pointed wings, and cuckoos with the longest primary number 4 and 5 have rounded wings.The details of the wing formula vary within cuckoo species, but this variation is much less than between groups of closely related species of cuckoos (Stephan 2001a). Plumage colors of specimens in different museums were compared with reference to a standard color guide (Kornerup and Wanscher 1967). Also useful were color photographs, taken with a standard

Plan of the book xxi gray reflectance card placed under the specimens. A Nikon-F3 camera was attached to a ring flash at 0.7 or 1.0 m from the specimen. Photographs taken under these conditions were compared, for birds in different museums. Photographs of birds such as the Little Bronze-cuckoo Chrysococcyx minutillus species complex show the glossy color to be the same across the visual field, except in the center of the body plumage where it appears dark. Glossy cuckoos with bronze gloss are more greenish than in their photographs, and these specimens were either compared directly with each other, or their photographs were compared with other photographs taken in the same conditions. ‘Sources’ indicates the museum collections that were used for species descriptions and, in some cases, the publications used for the appearance of nestlings. Colors of the bare skin on the face, the eye-ring, the iris, the bill and the feet were recorded from photographs and the labels on museum specimens. Where the colors of plumage and the bare parts differed between the sexes and ages, these are described. Subspecies are recognized when populations are nearly exclusive in size, color or other morphological characteristics, in different parts of their geographic range. Geographic variation in birds involves average size differences between birds in different areas, or a continuous variation between over areas. For example, Yellow-billed Cuckoos in western North America are larger on average than birds in eastern North America, but most birds are in the size range that overlaps between the eastern and western birds, and no subspecies are recognized in this case. Banders and collectors who see these birds in migration or in their winter range in South America, cannot know for certain the breeding areas where the birds originated in North America. Although early work with cuckoos involved collecting birds in the field and studying them in the museum, recent work in the field has involved conservation of their populations in nature. Wilson (1993) cites Baba Dioum:“In the end, we will conserve only what we love, we will love only what we understand . . .”. Conservationists are concerned about the birds whose numbers are low, and whose continued existence is threatened by human activity. This concern is expressed through the protection of

their habitats, to preserve the wild birds alive in natural conditions. Even though these biologists favor an agenda for conservation, we make our best estimates of the distinctiveness of bird species, on the basis of their biology rather than their conservation status. Conservationists have sometimes argued that if each population is given the special taxonomic status of a species, this will increase the legal and social protection of the population. There is, however, no evidence that systematic biologists have had a compelling impact on our own population growth, economics, migration, commercial logging or agricultural techniques. It is questionable whether biologists ought to set their scientific standards for a perceived social gain (Slobodkin 2001), such as in naming every population as a species or a subspecies, which practice becomes a political statement rather than a biological description. Even though we look for arguments and means by which to preserve wild birds and their habitats, the long-term outlook for these birds, and for tropical forests in general over the years to come, is not encouraging (Struhsaker 1997, Heaney and Regalado 1998, Oates 1999, Holmes 2000). Morphological descriptions are used to identify birds in the field or in a museum, and these descriptions are presented in the book. In earlier days a systematic revision provided a diagnosis, that is, a summary of characters that differentiate a taxon from related or similar taxa, and genera were recognized and diagnosed when an authority perceived “really trenchant external characters” (Peters 1940:45). In practice, the description of a species became the standard for the diagnosis of a genus (e.g., Cabanis and Heine 1863). Because similar body forms have evolved in different lineages of birds, as molecular genetic studies have shown (e.g., Sibley and Ahlquist 1990, Cibois et al. 1999, 2001, Sorenson and Payne 2001), the book emphasizes description rather than diagnosis. The book takes a new look at the relationships among the cuckoos, using phylogenetic analysis of discrete molecular genetic characters. The genetic characters (that is, the sequences of nucleotide bases in the DNA) determined in the study, can be accessed by computer in Genbank, the international register for comparative genetic data. These

xxii Plan of the book sequences were used to determine the groupings of cuckoo species into clades (lineages or branches of common ancestry) (Hillis et al. 1996). Molecular genetic analysis uses a large number of characters (the four nucleotide bases and their sequences at thousands of sites in the mitochondrial gene).These genetic characters provide a much larger data set than distinct morphological characters have ever allowed. Molecular genetic profiles were determined for each cuckoo species. The evolutionary relationships among cuckoos were estimated from the molecular genetic sequences and computer programs (PAUP*: Swofford 2001; POY, Gladstein and Wheeler 1996) which use parsimony, a principle of the smallest number of evolutionary changes that are necessary to derive the relationships from a common ancestor in a phylogeny. The phylogenetic estimates of relationships among the cuckoos in the molecular genetic results, were similar to the traditional sets of genera and subfamilies of cuckoos as recognized by morphologists and other museum systematists. Where they differed, the genetic phylogeny was taken as the best estimate of common ancestry, and these results are used throughout the book. I examined nearly all specimens from which the genetic samples were taken. A few re-identifications of these cuckoo specimens was a result (Payne and Sorenson 2003). The evolution of social behavior, such as cooperative breeding and brood parasitism, is a question of general significance. Phylogeny is used not only to clarify the evolutionary relationships among cuckoo species, but also to determine which behaviors evolved once or more than once, and the circumstances in the most closely related lineages that may have led to the evolution of these behaviors (Brooks and McLennan 1991, Edwards and Naeem 1993, Ligon 1993, Sorenson and Payne 2001, 2002). That is, cuckoos are a case study of evolutionary history and adaptation to behavioral ecology. Determining their phylogeny is a way to test the course of evolution of brood parasitism from a non-parasitic ancestor. These questions have been

addressed before (Poiani and Elgar 1994, Poiani 1998) using the estimate of cuckoo relationships in Sibley and Ahlquist (1990), yet that estimate was based on a genetic technique that did not use discrete characters, and it did not include all the major groups of cuckoos. The phylogenetic analysis of molecular genetic characters in the present study was designed not only to get a best estimate of relationships among the cuckoos, but also to determine processes in the evolution of social behavior and breeding biology. The arrangement of subfamilies, genera and species follows the phylogenetic convention of listing the basal taxa first, and the recent radiations or “crown” taxa last, in the order of their branching in the evolutionary tree (Wiley et al. 1991). This convention has long been used in some avian groups such as waterfowl, in contrast to recent reviews of cuckoos which first list the familiar cuckoos of Europe.Where the first convention did not apply, as in symmetric branches of the tree, other conventions were used. First, species with little or no geographic variation between populations (subspecies) were listed, before species with greater geographic variation. Second, where no subspecies were involved, the species with the more northern extent of distribution was listed first. Application of the “basal group first” norm caused the New World Cuckoos to be listed before the Old World cuckoos, and within the New World clade, the group-living crotophagine cuckoos are listed first. The order in which birds appear in the list does not necessarily mean that ancestral cuckoos were cooperative-group living birds. Within a species the geographic forms are listed from north to south and from east to west, with no necessary phylogenetic interpretation. Geographic names in the book are mainly those in the Times Atlas of 1993. In Indonesia the historical names of the Moluccas, now Maluku, and the Lesser Sunda islands, now Nusa Tenggara, have been retained, as they are widely used in current regional works on birds (White and Bruce 1986, Coates and Bishop 1997).

PART I General chapters

This page intentionally left blank

1 Introduction to the cuckoos

The parasitic breeding behavior of cuckoos has fascinated people for centuries. The brood-parasitic cuckoos lay their eggs in the nests of other kinds of birds, and never rear their own young ( Johnsgard 1997, Rothstein and Robinson 1998, Davies 2000). The natural history of the Common Cuckoo Cuculus canorus and the Great Spotted Cuckoo Clamator glandarius, the two species that live in Europe where they have been observed for many years, is well known. In Classic times,Aristotle studied the brood-parasitic behavior of Common Cuckoos (Friedmann 1964b,Aristotle 1991), and in Britain Shakespeare referred to cuckoos in his plays. The term “cuckoldry” was often used in the sense of an adulterous affair, and the implications of immorality in a later period led to a censored version of Shakespeare that eliminated the term “cuckold” (Bowdler 1861, Hamilton 1996). Many other species of brood-parasitic cuckoos live in Africa, Asia and Australia and in Central and South America where they have been watched by resident naturalists. Brood-parasitic cuckoos are diverse with 56 species in the Old World and three in the New World. Nevertheless, brood parasitism is only one of several breeding behaviors in this family of birds. A few cuckoos, the anis and the Guira Cuckoo of the New World, are sociable, with several pairs sharing a nest where they lay their eggs and care for the young; these are the cooperative breeders. Their cooperation is balanced by competition, as a female may remove the eggs of other females when she lays her own. In contrast to these birds which have gone to extremes of parental care and social behavior, most species of cuckoos live in solitary pairs and regularly build a nest and raise their own young.

Even these cuckoos occasionally lay their eggs in the nests of others, either a nest one of their own species or a nest of another kind of cuckoo. One cuckoo species is extinct, and eight cuckoo species are threatened or endangered at the global level. The distribution of cuckoos is cosmopolitan, and most species live in the Old World tropics.The variation in social behavior and parental care of cuckoos may be unmatched among the bird families of the world. The cuckoos are the most successful broodparasitic birds. They have the largest number of species, the largest number of host species that rear their young, and worldwide distribution.The cuckoos are one of the five families of birds in which brood parasitism is the only life style for one or more species—the others are two families of songbirds (the Old World finches and the New World cowbirds), the honeyguides, and the ducks. One duck is a dedicated brood parasite; indeed other waterfowl sometimes lay their eggs in the nests of other ducks of the same species or a different species. In each of these other families, an obligatory breeding style of brood parasitism has evolved only once (Sorenson and Payne 2001, 2002). Brood parasitism has originated more than once in cuckoos. Results of molecular analysis in this study indicate that the parasitic lineages of cuckoos are not all each others’ closest relatives. Because brood parasitism has evolved more than once, the cuckoos provide a replicate test of the behavioral context of the course of evolutionary changes when parental care was transformed into brood parasitism. Young cuckoos are altricial. They remain in the nest where they depend on their parents or foster

4 The Cuckoos parents to provide food, and they grow rapidly.The young of nesting cuckoos leave the nest days before they are able to fly, while the young of broodparasitic cuckoos take longer to fledge and are welldeveloped when they leave the nest. Nesting cuckoos in the New World have the shortest incubation periods and the shortest nestling periods of any birds. Even the Old World coucals have a short nestling life, unusual for such large birds.The variable period in which the cuckoos depend upon parental care is closely associated with the life style of nesting and brood parasitism. The mating systems and social organization of cuckoos range from monogamous pairs that stay together in a social bond, to the cooperatively breeding anis and Guira Cuckoo where incubation and nestling care are shared by group members that use a single communal nest to rear their young. Among these birds there is genetic evidence of polygyny and polyandry as well as social monogamy. Cuckoos have a long evolutionary history. Studies in molecular genetics have estimated that cuckoos occurred as a distinct lineage more than 60 million years ago, a lineage as ancient as that of any modern bird (Sibley and Ahlquist 1990). Fossil cuckoos have been known for only a part of this time. These fossil birds are as varied as modern cuckoos, and some appear to have been flightless. Cuckoos occur throughout the world on all habitable continents and on the island areas of southern Asia, and they are an important part of the biological diversity of tropical birds. Much of what is known about cuckoos came from birds collected in the 1800s when they were first discovered by western Europeans. A few cuckoos were described by their collector and discoverer, Stamford Raffles (1781–1826), natural historian, ethnologist, Governor of Java and founder of the trading station that became Singapore (Raffles 1817, 1830). He began his work as a clerk in London and read widely, mastered languages, cultivated friends and gained a position in Penang. Based on his interests and his rewarding discussions with the Viceroy of India, he was able to study the area’s history, culture and wildlife (Raffles 1830, Mearns and Mearns 1998). Raffles was unusual among early collectors for his access to the publishing world. Later collec-

tors visited the same areas and prepared specimens that made their way back to natural history museums, and in some regions resident naturalists were able to observe the details of their behavior and breeding biology. These collectors were employed by governments and private collectors, and most of their specimens were sent to natural history museums in Europe where the species were described and named (e.g. Blyth 1842, Gould 1845b, Cabanis and Heine 1863, Salvadori 1881, Shelley 1891). Linnaeus described more kinds of cuckoos than any other taxonomist, and half of the 22 species that he described are currently recognized as distinct kinds of birds (some others included a second sex), although the specimens from which they were described are no longer in existence. Tommaso Salvadori appears to hold the record for the number of new cuckoos described by a post-Linnaean ornithologist. Museum patrons employed bird collectors who visited Borneo, the Moluccas and New Guinea, and from specimens that made it safely back to Europe, Salvadori described 11 kinds of cuckoos which are now recognized as species and subspecies (Salvadori 1874, 1875, 1876, 1878a,b, 1879, 1881, 1889). He was based at the museum in Turin (Elter 1986,Violani et al. 1997), and his many type specimens in Turin, Genoa, Leiden and other museums rank him highest among his ornithological generation (Payne 1997b). He usually did not designate a single specimen as the holotype of a species, but rather syntypes of a type series, many of which went to other museums as “duplicates”, with some still being discovered in these collections. Salvadori worked with other families of birds as well. He worked for two years at the National Museum in London where he wrote monographs on parrots and pigeons, and he lived in Leiden where many of his type series are located (van den Hoek Ostende et al. 1997). In other museums in Germany, Cabanis and Heine (1863) described nine cuckoos which are now recognized as species and subspecies, and they described several cuckoo genera as well. A few species of cuckoos are known only from museum collections. The nests of several cuckoos have never been found, and the life histories of these and many other cuckoos are unknown. Many cuckoos live in the Old World tropics where few

Introduction to the cuckoos 5 observations of their life histories have been made in the forests, where these birds are hard to see as the nests are well hidden. Once a nest is found a predator is also likely to find it. Little collecting of anatomical and genetic materials for scientific research has been done in recent years, and the careful field studies that were the fieldmark of resident ornithologists have waned with changes in land ownership. Current research efforts have been limited by nationalistic zeal, and by the difficulty of productive research by unacclimated field biologists in the malarial tropics.The museum specimens that have been collected over the past two centuries are a valuable and unreplaceable resource and are our primary record of biological diversity.We continue to learn new details from these museum specimens, such as geographic variation and the plumages and molts of birds, and we can now retrieve genetic information from feathers or skins of these specimens. Field studies on behavior and recordings of the songs of cuckoos are much needed to understand these birds in natural conditions, and great opportunities remain to discover the life styles and the changing distribution of birds such as the cuckoos.

What are the cuckoos? Cuckoos are zygodactyl birds with the inner and outer toes directed backward and the other two toes directed forward. The bill is usually slender and slightly arched.The plumage of most cuckoos is soft and lax. The body form varies among species with their systematic relationships, body size and life style. Most arboreal cuckoos have a slender body and long tail, whereas terrestrial cuckoos are heavybodied and have long tarsi for their size (weight). The nostrils are round in many cuckoos, but slit-like in the coucals, in the crested cuckoos Clamator and Thick-billed Cuckoo Pachycoccyx audeberti, and to a lesser extent in Long-tailed Cuckoo Urodynamis taitensis, Common Koel Eudynamys scolopacea and Dwarf Koel Microdynamis parva, and slit-like in some malkohas also. Many cuckoos have long eye-lashes. The bill has no cere; the tarsi are scutellate and often unfeathered. The base of the tail has a naked and bilobed oil gland.

Cuckoos have a wing with 10 primaries and 9 to 13 secondaries (9 in several malkohas, in the large Caribbean species of Coccyzus (“Hyetornis”and “Saurothera” ), and in some squirrel cuckoos Piaya; 9 or 10 in Cuculus cuckoos; 10 in hawk-cuckoos Hierococcyx,Thick-billed Cuckoo Pachycoccyx,African Crested Cuckoo Clamator levaillantii and in some coucals; 11 in koels Microdynamis and Eudynamys, in some couas and coucals and in the Great Spotted Cuckoo Clamator glandarius); 12 in some other coucals, and 12 or 13 in the Channel-billed Cuckoo Scythrops novaehollandiae (Stephan 2001a).The alula is small in most cuckoos but large in the New World ground-cuckoos Neomorphus and Carpococcyx, and in the brood-parasitic American Striped Cuckoo Tapera naevia. In the last two cuckoos the large alulae are extended when the wings are spread in display.The wing is eutaxic.The pattern of wing molt is peculiar to cuckoos, where a wave of molt jumps over the neighboring old feathers. The odd-numbered primaries first drop and grow, followed by the evennumbered primaries. Cuckoos are capable of flight, varying from the swift and direct flight of the longwinged, long-distance migrant Common Cuckoo Cuculus canorus and Black-billed Cuckoo Coccyzus erythropthalmus of northern temperate regions, to the gliding flight of short-winged tropical forest-living malkohas and the slow awkward flight of anis, couas and coucals. The tail usually has 10 rectrices, though only 8 in anis Crotophaga and Guira Cuckoo Guira guira. The tail of some cuckoos is shorter than the wing, particularly that of the glossy cuckoos Chrysococcyx. The tail of other cuckoos is one and a half times the length of the wing, as in the arboreal squirrel cuckoos Piaya cayana and P. melanogaster, the yellowbills Ceuthmochares and several Asian malkohas, and the terrestrial ground-cuckoos Geococcyx and Neomorphus (Stephan 2001b). In North and Central America where all cuckoos have long tails, the area of the tail is larger in cuckoos than in any other family of bird of comparable body size (Hartman 1961). In many cuckoos the tail is strongly graduated, with the central feathers longest and the outer feathers short, and the steps marked by the conspicuous white tips of the tail feathers. In others the tail is rounded, square, or even forked, with the central

6 The Cuckoos feathers shorter than the outer feathers, especially in the Fork-tailed Drongo-cuckoo Surniculus dicruroides and Moluccan Drongo-cuckoo S. musschenbroeki. Cuckoos span a wide range in body size, with most species 16–70 cm in overall length.They vary in mass from 17 g with the smallest, Little Bronzecuckoo of Australasia, to 400 g with Giant Coua Coua gigas of Madagascar, and 700 g with the largest broodparasitic bird, Channel-billed Cuckoo Scythrops novaehollandiae of Australia (the next largest brood parasite is Black-headed Duck Heteronetta atricapilla at 513–565 g, Carboneras 1992). The largest nesting cuckoos are coucals in the New Guinea region, where Greater Black Coucal Centropus menbeki and Violaceous Coucal C. violaceus are more than 500 g, Goliath Coucal C. goliath are over 600 g, and Buffheaded Coucal C. milo are as large as 770 g.

The major groups of cuckoos The major groups of cuckoos include two groups with a common ancestry in the New World, two groups in the Old World, and one group that occurs in both regions. In the past, some of these groups have been called families, and others have only recently been recognized as lineages with a common evolutionary history, especially the Cuculinae, which include both New World and Old World cuckoos, and both nesting species and brood-parasitic species.

1. Crotophaginae The crotophagines are groupliving cuckoos and all cuckoos with this life style live in the New World.They have robust legs and a long tail, streaked brown plumage in the South American Guira Cuckoo Guira guira and black plumage and a deep compressed bill in the three ani species Crotophaga. Living in aggressive and noisy social groups, several pairs defend a common territory and share a joint communal nest where two or more females lay their eggs (in anis, some nests may be of single females) and the adults all rear their common young together as cooperative breeders. The nest is flat or a shallow bowl, built in a tree.The nestling mouth is pink with white marks on the palate and tongue. Anis often have a pungent odor, noticed when birds are held in the hand or when they are at their night roost. The social or adaptive

significance of the smell is unknown. The crotophagines have large and unique anal glands, but whether these produce chemical deterrents to predators or chemical signals to social partners or are accessory reproductive structures is unknown (Quay 1967).The ridge on the bill is shaped by an underlying ridge in the nasomaxillary skeleton, especially in Greater Ani Crotophaga major.Awkward in movements with the wings and tail seemingly disconnected from the body, anis are unlike the graceful arboreal cuculine cuckoos.Anis are slow and clumsy as they flutter and clamber through low bushes and weeds, where their flopping wings and waving tails flush out insects in dense vegetation; Guira Cuckoos fly like kites in the wind.

2. Neomorphinae Ground-cuckoos of the New World have short wings, long legs and a long tail. They include a dry scrub-forest terrestrial cuckoo of Central America (Lesser Ground-cuckoo Morococcyx erythropygus), lizard-catching roadrunners (Geococcyx) of the semi-arid regions of North America, and large South American ground-cuckoos Neomorphus that follow ant swarms in the tropical forests and eat the insects that escape the marauding ants. Roadrunners run across the ground at speeds of up to 30 km/hr in chase of their lizard and grasshopper prey, the birds holding their heads and tails level with the ground and swinging their tails like rudders. They often course along roads, paths and dry stream beds, and walk and run on their daily rounds for food and to patrol their territories. These ground-cuckoos can fly, though when they survey their area from an elevated perch they stay within a hop and a flap of the ground. The nesting ground-cuckoos build a nest and rear their own young nestlings which have conspicuous colored spots inside their mouths. The ground-cuckoos also include three species that are brood-parasitic (two species of Dromococcyx, both shy and secretive, and the American Striped Cuckoo Tapera naevia), and are forest dwellers.They parasitize several species of passerine hosts that breed in covered or domed nests, and occasional other hosts that breed in open nests. 3. Centropodinae Coucals are large groundforaging birds with long stout feet, a long straight

Introduction to the cuckoos 7 claw on the inner hind toe in most species (the claw is the “foot spur” of the genus Centropus), short rounded wings and a long broad tail. The plumage is black, rufous brown, or white, or a combination of these colors; the plumage is shaggy and in many species the neck and breast have spiny hackles. Coucals live in the Old World, mainly in the tropics from Africa and Asia to Australia, New Guinea and the Solomon Islands, and on many islands in Wallacea and the Papuan region. Coucals live in forest, swamp and marsh habitat, and spend most of their time on the ground, and in dense marshes they are most often seen in their slow and awkward flight. Most coucals build large globular or domed nests of grass and leaves with a side entrance. Most make deep resonant cooing calls, some species in a “water-bubbling” pattern which recalls the gurgling sound of water as it rushes from a narrow-necked bottle, and some species call in duets. Nestling coucals have an odd set of stiff natal down (trichoptiles) and as the body feathers grow these hair-like structures are attached to the tips of growing feathers, then fall off when the birds fledge; and they have conspicuous raised colored spots on the palate.Adult coucals have a pheasant-like body form with long, robust legs and a long tail, coarse plumage with stiff hackles, and a bronchial syrinx. Males of some coucals have a single testis while others have two testes; the number can vary between one and two within a species and a local population (Mayr 1937, Mayr and Rand 1937, Rand 1942a).

4. Couinae These are large colorful groundcuckoos of Madagascar and southeast Asia and have long tarsi. Ground-cuckoos Carpococcyx of tropical forests in Asia have plumages of dove-like greens and blues, bright bare skin around the face and long eye lashes, and they build flat platform nests. Couas Coua are large cuckoos in Madagascar. They have soft and lax plumage with dove-like colors of pastel pinks, purples and peach (blue in the Blue Coua Coua caerulea), a long tail and large feet, colorful bare areas of skin around the eye, and long eye-lashes. Most are terrestrial. The arboreal Blue Coua has long legs like the terrestrial couas and like them it walks, but in the trees.The Red-capped Coua Coua ruficeps is mainly terrestrial and walks on the ground,

and along branches when it is in a tree (Berger 1960). One species of coua, a snail-eating specialist on Ile Sainte Marie northeast of Madagascar, was last seen around 1835. This is the only cuckoo species known to have become extinct in the past two centuries. Nestling couas and Asian groundcuckoos lack natal down, and the palate and tongue have conspicuous patterns of raised white or blue spots that contrast with the red mouth lining and may be a signal to their parents during parental care.

5. Cuculinae These include the long-tailed arboreal nesting cuckoos of the Old World and New World.The Old World nesting birds in this group are known as the malkohas and are now grouped into six genera in Asia: Rhinortha, Taccocua, Zanclostomus, Phaenicophaeus, Dasylophus and Rhamphococcyx, and a seventh, Ceuthmochares, is found in Africa.“Malkoha” is a Sinhalese name in Sri Lanka (Ceylon) for the Red-faced Malkoha Phaenicophaeus pyrrhocephalus. It has been widely used for other species in the Oriental Region (Levaillant 1806, Raffles 1822, Jerdon 1862) and as a generic name Malcoha Schinz 1821. Malkohas have strong, unfeathered tarsi, short rounded wings, and long graduated tails tipped with bold spots and bars. Malkohas are skulking arboreal cuckoos of forests and thickets. Many have brightly colored bare skin around the eye and face, and the bill is large, often arched and brightly colored. The German term for the birds, “Buntschnabelkuckucken,” translates as the descriptive “colorfulbilled cuckoos.” The wing feathers often have wide vanes and the body feathers have stiff, shiny shafts in a color that contrasts with the rest of the feather. Malkohas build shallow nests, lay chalky white eggs and rear their own young. The malkohas are brightly-colored in plumage, with swollen and brightly colored bills, short wings and long tails. All are arboreal except for Sirkeer Malkoha Taccocua leschenaultii which is terrestrial. A cultural awareness of cuckoos as brood parasites in Europe caused some ornithologists to guess that the other cuckoos were also brood parasites, and Newton (1896) suggested that malkohas were brood-parasitic, but field ornithologists in India and the Malay Peninsula, Borneo and Sulawesi had by that time observed adult malkohas at their own nests. Little is known

8 The Cuckoos about the details of nesting such as the incubation period, nestling period and parental behavior in these birds. The brood-parasitic Old World crested cuckoos Clamator appear to be closely related to the malkohas. Several cuckoos of the New World are the ecological counterparts of malkohas and are closely related to them. Most are arboreal insect-eaters and have long tails. The coccyzine cuckoos including Coccyzus have long wings, and several species are long-distance migrants between continental North and South America, or between temperate South America where they breed and tropical South America where they winter. Others are large-bodied cuckoos that live on islands in the Caribbean and feed on lizards. The New World cuckoos build saucer-shaped nests, lay white or blue eggs, and rear their own young. The young have raised papillae inside the mouth; the color of the papillae contrasts with the palate. Cuculines also include the brood-parasitic cuckoos of the Old World, the tribe Cuculini. Many are migratory with long, narrow, pointed wings, long tails, and short legs with the tarsus feathered at the base. The tail is often graduated with the central feathers longer than the others. As noted by Aristotle, Common Cuckoos Cuculus canorus look like hawks (notably Merlin Falco

columbarius) in plumage, shape and flight; and now we know the hawk-cuckoos Hierococcyx have rounded wings like bird-hawks (Accipiter). Most brood-parasitic cuckoos are dull in plumage, and their cryptic appearance may help them to avoid detection by their hosts (Payne 1967). The parasitic glossy cuckoos Chrysococcyx of Africa and Australasia, however, are conspicuous in their colorful plumage. They are small and at first glance could pass for insect-eating songbirds, and sometimes they even feed in mixed-species flocks with songbirds. Old World brood-parasitic cuckoos lay their eggs in the nests of many passerine host species. Their eggs vary from blue and green to chocolate brown to white, spotted or unspotted, and some species of Old World brood-parasitic cuckoos have more than one kind of egg marked by different colors and patterns. The nestlings are naked when they hatch (young Shining Bronzecuckoos Chrysococcyx lucidus and Little Bronzecuckoos have sparse natal down). Inside the mouth the young lacks visual effects except for a healthy color of red, orange or yellow.The young of one brood-parasite, African Thick-billed Cuckoo, however, have a bright orange palate with contrasting pale spots, and nestling Rufous Hawk-cuckoo Heirococcyx hyperythrus have a palate of yellow and pink.

2 Distribution, habitats and conservation status

Distribution Cuckoos occur around the world in tropical and temperate regions. More species live in the Old World (n ⫽ 109) than in the New World (n ⫽ 32). The northernmost cuckoos (Common Cuckoo Cuculus canorus, Oriental Cuckoo C. optatus, Indian Cuckoo C. micropterus,Asian Lesser Cuckoo C. poliocephalus) occur in northern Europe and Russia in summer.They are migrant brood-parasites that lay in the nests of migrant Palearctic songbirds and then winter in the tropics, sometimes crossing the equator and even passing over the Indian Ocean from Asia to Africa. In the New World the two species (Yellowbilled Cuckoo, Black-billed Cuckoo Coccyzus erythropthalmus) that breed in the United States and Canada are mainly warm-weather birds that spend most of the year in the New World tropics. No cuckoo species occurs regularly and breeds in both the Old World and the New World, although vagrants from northeastern Asia (Common Cuckoo and Oriental Cuckoo) turn up in the Alaskan archipelago and northern Alaska in migration season, and vagrant Black-billed Cuckoos and Yellow-billed Cuckoos from northeastern North America show up in many years in Europe from Iceland to Italy. Cuckoos are widespread as breeding birds on islands, especially in the Indonesian archipelago and the New Guinea region. Migrant cuckoos occur on these islands in winter, mainly species from the northern hemisphere; Australian cuckoos also live on the islands during the austral winter. Cuckoos are not as likely to turn up on oceanic islands as are fruit-eating pigeons and doves, nectar-feeding white-eyes or the rails, yet migrant Oriental

Cuckoos from northern Asia occur as far from their breeding grounds as Palau in the western Pacific and Australia in the Southern Ocean. Perhaps the most widely-traveled species is Long-tailed Cuckoo Urodynamis taitensis which migrates from its breeding area in New Zealand to winter on islands in Melanesia, Micronesia and Polynesia across a wide area of the tropical Pacific. Three species (Common Cuckoo, Asian Lesser Cuckoo, Jacobin Cuckoo Clamator jacobinus) migrate from Asia to Africa for the non-breeding season, and some birds appear during migration season on islands in the Indian Ocean. Cuckoos are unusual birds in this India-Africa migration pattern. All major clades of cuckoos occur in South America, Africa, and southern Asia through to Australia.The biogeography of cuckoos suggests an ancient origin and radiation of cuckoos in the southern region of the world. Geologically this area was the ancient tectonic plate Gondwanaland. Cuckoos in the northern region are all closely related to cuckoos in the southern region and appear to have derived from the southern cuckoos. The ancient diversity of cuckoos that is suggested by the genetic differences between major groups (Sibley and Ahlquist 1990) point to an early origin of the major clades, perhaps before the Eocene and at the time of the breakup of the Gondwana plate; yet there is no independent calibration of this date from dated fossils. The North American Coccyzus (including the large lizard-cuckoos “Saurothera” and “Hyetornis” of the West Indies) is closely related to the Old World malkohas and appears to have been derived from them, perhaps following the dispersal of ancestral nesting cuculine cuckoos by way of

10 The Cuckoos South America.Within the Old World the distributions of certain cuckoo groups suggested to some workers, an initial separation traced to plate tectonic events. Marchant (1972) suggested this origin for the biogeographic separation of Chrysococcyx glossycuckoos of Australasia and Africa, and Cracraft (1988) suggested the same for the Carpococcyx groundcuckoo species in southeast Asia, Borneo and Sumatra. So far there is no evidence that the genetic distances observed between the clades and species of cuckoos correspond with the genetic distances expected from these tectonic events, based on mutation rates and the length of time since the events. The goal of biogeographic models should be to test whether the geological and the genetic times of divergence coincide (Zink et al. 2000).

Habitats Cuckoos are mainly arboreal birds of forest and woodland. Their habitats are varied but include many primary, more or less undisturbed tropical rain forests. Other cuckoos live in secondary forest and some of these have adapted to life in tree plantations, particularly that of cocoa where larger trees provide shade for the cocoa plants.The forests they inhabit range from primary evergreen forests to semi-arid scrub in Madagascar. Cuckoos often occur in the canopy and on dense vines and rattan in Asia, and the ground cuckoos are mainly forest birds. A few cuckoos live in coastal mangroves (Mangrove Cuckoo Coccyzus minor, some malkohas and coucals in southern Asia, and some broodparasitic cuckoos in Australasia). Cuckoos also live in open brushy country in Australia and the plateau country of Africa, in the southwestern United States, Mexico and the drier parts of Central America, and in southern South America. The diversity of cuckoo behavior is highest in the New World tropics. Here, many cuckoo species nest as pairs and rear their own young, others are cooperative breeders that care for the young of others in their social group, and three are brood parasites which do not rear their own young. Cuckoos are especially numerous and diverse in Amazonian Brazil. Near Manaus there are seven species of cuckoos, only one a brood parasite, one a cooperative breeder, and

others solitary birds, or ones that join mixed-feeding flocks but nest in pairs away from other birds (CohnHaft et al. 1997). In one of the largest forest reserves in the world,Tapajós National Park, with 448 species of birds, seven are cuckoos (Oren and Parker 1997); in Jaú NP, with 445 species, nine are cuckoos (Borges et al. 2001); on the Middle Rio Jiparaná, Rondônia, with 459 species, 12 are cuckoos; and in the Alta Floresta region in extreme northern Mato Grosso, with 474 species, nine are cuckoos including a nonbreeding austral migrant (Zimmer et al. 1997). In all these areas most cuckoos are of the solitary nesting species. The Atlantic coastal forests in Brazil have as many as ten cuckoos, although one, a subspecies of ground-cuckoo, may now be extinct in that region (Parker and Goerck 1997). The number of cuckoo species is highest in forests of the Old World tropics. In southern Asia at a local scale of a 2 km2 area in the Malay Peninsula, Khaso Nor Chuchi has 247 breeding bird species with 29 resident cuckoos (Round and Treesucon 1997), the Pasoh Forest Reserve has 188 species of birds and the Kerau Wildlife Reserve at Kuala Lompat has 195 species, each site within the geographic range of 21 cuckoo species (Wells 1999).The Tekam Forest Reserve has 225 species with 12 resident cuckoo species, ( Johns 1989). In the Vu Quang Nature Reserve,Vietnam, there are 220 resident bird species with 9 cuckoos, and in 14 other sites in Vietnam there are 16 cuckoo species (Eames et al. 2001). On a larger scale, the Malay Peninsula has 686 species with 22 cuckoos that are unlikely to breed there (Wells 1999) and Sabah has 339 resident species with 21 cuckoos (Sheldon et al. 2001). Island areas are less species-rich both in birds on the whole, and in cuckoos. In the Lesser Sunda Islands, 251 species are known on Flores, nine of them cuckoos (Verhoeye and Holmes 1998); and 182 species on Sumba, six being cuckoos (Linsley et al. 1998). In Central Africa, as many as 16 resident species of cuckoos live together in forests and on forest edges. Four are coucals Centropus species, one is a malkoha, Chattering Yellowbill Ceuthmochares aereus, and the others are brood parasites that lay in nests of forest passerines. In Cameroon in Lobéké FR where 305 species of birds are known there are 13 cuckoos, 10 of them in forest habitat; and Mt. Kupe has 324 resident

Distribution, habitats and conservation status 11 species with 10 cuckoos in Congo-Brazzaville Nouabalé-Ndoki has 273 species with 12 cuckoos, Odzala NP has 435 species with 17 cuckoos, and the Kouilou basin has 425 species with 16 cuckoos (Dowsett and Dowsett-Lemaire 1997, DowsettLemaire and Dowsett 2000, Bowden 2001), in the Central African Republic Dzanga-Sangha has 360 species with 11 cuckoos (Green and Carroll 1991, P. Beresford in litt., AMNH) and in Uganda Kibale Forest has 410 species with 13 cuckoos (Struhsaker 1997). In each of these Old World areas, more than half the cuckoo species are brood-parasitic. The difference in the number of cuckoo species between the New World and Old World lies mainly in the greater number of brood parasites in the Old World, and this is a result of different evolutionary histories, rather than any obvious ecological difference between the two regions. Cuckoos live not only in the warm tropics but in cool and wet conditions as well, in forests and in more open habitats. When their plumage becomes wet, the cuckoos use the sun to dry and these birds are best seen when they expose themselves high on a perch in the early morning or after a shower, spreading their wings and tail and lifting their back feathers, exposing the skin to the sun. Sunning behavior is well known in coucals and glossy parasitic cuckoos in the Old World, and in Coccyzus cuckoos, roadrunners Geococcyx californianus and anis Crotophaga and Guira in the New World. Ani plumage becomes soaked as the birds forage in wet grass and herbaceous vegetation. Guira Cuckoos Guira guira have heavily pigmented black skin between the dorsal feather tracts, and in the morning on cold days the birds droop their wings, turn their back and expose the black skin toward the sun to dry the plumage and warm the body. Anis adapt to cool weather and to poor food conditions by becoming semi-torpid; they drop their body temperature a few degrees at night, and when their body temperature drops the anis can still fly after two days of fasting (Warren 1960). A few cuckoos live in seasonally hot, arid and semi-arid habitats. These include roadrunners in the New World, a few couas Coua in Madagascar, and several brood parasitic cuckoos in Australia.

While the dry-country brood parasites escape the extreme conditions when they migrate, the nesting cuckoos typically live in the deserts as residents. Cuckoos have adapted well to human changes in habitats in some areas. In the New World tropics, anis Crotophaga have spread with forest clearings. In southeast Asia the koels Eudynamys scolopacea have spread to new areas and new host species that do well in towns and cities. In Australia the extensive plantings of gardens and flowering trees have attracted nectar-feeding birds. In southeastern Australia, Common Koels and Channel-billed Cuckoos have increased in numbers in the past ten to twenty years along with their host species, and in southwestern Western Australia Pallid Cuckoo Cacomantis pallidus are common where the large honeyeaters especially Red Wattlebird Anthochaera carunculata live in parks and gardens. Certain ground-living cuckoos including Gabon Coucal Centropus anselli, Red-billed Ground-cuckoo Carpococcyx renauldi and Red-capped Coua Coua ruficeps are often seen when they visit human areas and feed on camp refuse such as rice and noodles and on the rich insect life downstream from camp latrines.

Conservation status Although cuckoos are at some risk wherever their habitat is shared with humans, most cuckoos are not at immediate risk of extinction for at least another generation of humans. Unless otherwise indicated in the species accounts, the cuckoos are not considered threatened, when judged by the criteria in Collar et al. (1994) and BirdLife International (2001). The cuckoos most at risk are forest cuckoos. Tropical evergreen forests, the unique achievement of millions of years of nature, are being destroyed by logging for widescale economic consumption, and human populations are large and growing and are removing forest to cultivate open lands. As their forests are logged, many cuckoos are unable to adapt to any remaining scrub and secondary forest that might survive. Furthermore, forests once logged are open and vulnerable to fires and encroachment. In the past two decades, fires have burned millions of

12 The Cuckoos hectares of logged forests in Borneo and Sumatra (MacKinnon et al. 1996, Curran et al. 1999, Francis 2001). Global-imaging satellite records of change in humid forests of tropical regions between 1990 and 1997 have determined an annual deforestation rate of 1–4% in southeast Asia, 1–5% in Africa and Madagascar, and 3–4% in the upper Amazon (Achard et al. 2002). In the Old World the cuckoos at risk include the Red-faced Malkoha Phaenicophaeus pyrrhocephalus, Sumatran Ground-Cuckoo Carpococcyx viridis, Black-hooded Coucal Centropus steerii, Javan Coucal C. nigrorufus and Green-billed Coucal C. chlororhynchos. In the New World there are questions about the continued existence of Cocos Cuckoo Coccyzus ferrugineus, Rufous-breasted Cuckoo C. rufigularis and Banded Ground-cuckoo Neomorphus radiolosus (Collar et al. 1994, BirdLife International 2001). Several other species and subspecies are considered near-threatened (Collar et al. 1994), especially birds whose distributional range is small and limited to islands (Coral-billed Ground-cuckoo Carpococcyx renauldi,Verreaux’s Coua Coua verreauxi, Kai Islands race of Pheasant-coucal Centropus phasianinus spilopterus, Biak Coucal C. chalybeus, Short-toed Coucal C. rectunguis, Rufous Coucal C. unirufus and Scaled Ground-cuckoo Neomorphus geoffroyi squamiger. Although no brood-parasitic cuckoos appear on the lists of threatened and endangered birds, the loss of forests will have an impact on these birds, as well as on their host species. Most critically, the lowland and montane forests in the Malesian tropics of Malaysia, Indonesia and the Philippines are at risk, and with loss of forests these areas and the world are in danger of losing their cuckoos and the rest of their birds (Whitten et al. 1987, 1996, Collins et al. 1991, Sayer et al. 1992, Heaney and Regalado 1998, Kennedy et al. 2000, BirdLife International 2001). At least one brood-parasitic cuckoo should be added to the list of species at risk, Philippine Drongo-cuckoo Surniculus velutinus. Formerly regarded as a subspecies of a single species of drongo-cuckoo, the drongo-cuckoos are now known as four distinct species on the basis of differences in songs and in their plumage. More than 90% of forest habitat in the Philippines has been lost in recent years.

Preservation of their forest habitats is the only way to maintain the numbers and diversity of birds: necessary but perhaps not sufficient if hunting and trapping of birds continue. In addition to forest cuckoos, the cuckoos of open woodlands have seriously decreased in numbers. In Brazil, in the interior part of São Paulo State the number of species reported by earlier observers but no longer seen, is much greater than the number of species that have disappeared in the humid coastal areas, where the remaining forests are protected by law. Even there, birds are hunted and captured for the cage bird trade, and habitats of bamboo and coastal flats are developed for human housing. In the mesic and drier inland region of São Paulo, the open woodland savannas (“cerrado”) have been destroyed for large ranches, replaced by sugar cane for alcohol production, and coffee and oranges for export crops, and now the small dry forest and savanna fragments cover less than 1% of their original area. In inland regions of Brazil, Pearl-breasted Cuckoo Coccyzus euleri have little remnant habitat and their numbers are much reduced (Willis and Oniki 1992). In North America, Greater Roadrunners were controlled with a federal or state bounty on their heads as predators of quail, early in the twentieth century. In fact, these cuckoos hardly ever take quail, whereas they often take crop and household pests. Even though roadrunners have a traditional cultural status in the North American southwest (Dobie 1939), they are often shot as targets and trophies. Large ground-cuckoos undergo local extinction when their small forest patches support only a few birds, and the destruction of nearby habitats results in the loss of opportunity for birds in neighboring areas to disperse into the remnant forests. Local extinction has been documented on Barro Colorado Island, Panama, a forested hilltop that between 1912 and 1914 was isolated from the mainland forest when the Chagres River was dammed to form the Panama Canal.The area of the island is 1564 hectares, or six square miles.The site was designated a biological reserve in 1923 and has been the focus of long-term behavioral and ecological studies of birds and other tropical life (Eisenmann 1952, Willis 1979, Leigh 1999). When

Distribution, habitats and conservation status 13 the nearest populations of resident sedentary birds were lost, as forest habitats were cut around Gatun Lake, no source populations remained to repopulate the island. Rufous-Vented Ground-cuckoo Neomorphus geoffroyi was the first bird to become extinct on the island; it was last seen in 1935. The brood-parasitic Pheasant Cuckoo Dromococcyx phasianellus, was the second cuckoo lost; it was last seen on the island in 1971 as its shrub habitat was overgrown by forest (Willis and Eisenmann 1979). Of special interest and concern is Sumatran Ground-cuckoo Carpococcyx viridis, a fruit-eating bird that had not been seen by western ornithologists since 1916, but was rediscovered when one was captured in a live trap, photographed and released in 1997; another one was seen at another site in 2000 (Zetra et al. 2002). Sumatran Ground-cuckoo may persist in dipterocarp forests, where vulnerability to snaring and hunting, the widespread failure of these forests to fruit in certain years ( Janzen 1974, Whitten et al. 1987), and the loss and fragmentation of forests in response to logging and burning which restrict the dispersal of large vertebrates (Curran et al. 1999, Curran and Leighton 2000), all threaten the survival of large forest animals including the Sumatran Ground-cuckoo. The only cuckoo that is known to have become extinct as a species in historical times is Delalande’s or Snail-eating Coua Coua delalandei, which disappeared from Madagascar around 1835 during a time of deforestation. Hunting may have been involved in its extinction, along with the introduction of rats which depleted the terrestrial snails that were the coucals’ food. In regions where forest habitats have been destroyed, the species that do well in disturbed habitats have become more common and the forest species have become rare.This has led to some concern that replacement of the rare species is a result of competitive exclusion by the more common species. For example, the rare Black-hooded Coucal Centropus steerii and the more widespread Philippine Coucal C. viridis live in cut-over former forested lands in Mindanao. Elsewhere in southeast Asia and the Sunda Archipelago the uncommon forest-living Short-toed Coucal C. rectunguis, Javan Coucal C. nigrorufus and Green-billed Coucal

C. chlororhynchos are decreasing and their place taken by the widespread Greater Coucal C. sinensis. Disappearance of these rare coucals results from destruction of their habitat. It is unlikely the rare coucals would recover their numbers if the more common coucals were controlled. Conservation efforts are better applied to the protection of a diversity of habitats that are suitable for these rare coucals (BirdLife International 2001). Humans have used cuckoos for food and traditional medicine, (Payne 1997b). The cultural practice of folk medicine has declined and appears to be disappearing in rural areas, as in the use of the rare Green-billed Coucal in Sri Lanka (Wijesinghe 1999) and anis in the West Indies. Host species populations which the broodparasitic cuckoos parasitize are not a risk to cuckoos and their brood parasitism is not a conservation concern.When their hosts are thin on the ground, their nesting birds may be too few to support the cuckoos, because a population of brood parasites needs a large number of nesting hosts in order to maintain their own numbers from one generation to the next. Certain cuckoo species concentrate on only one or a few host species, and the population numbers of cuckoos are likely to be lower than the numbers of their hosts. Deteriorating habitat conditions for cuckoo hosts would drive the cuckoos into local extinction well before the hosts become threatened or endangered owing to cuckoo parasitism. In this sense the brood-parasitic cuckoos may be indicators of the environmental health of an Old World forest. Cuckoos depend on a healthy insect population for food and survival, and the brood-parasitic cuckoos depend on their hosts finding enough food to rear the cuckoo young. Environmental disturbances that lead to the reduction of hosts also lead to the loss of brood-parasitic cuckoos, and disturbances that reduce the area of wooded habitats and their insects lead to the decrease of other cuckoos as well. Cuckoos are among the first birds to disappear when a habitat is polluted by industrial and chemical waste. Their caterpillar prey accumulate toxins from chemical pollution. In North America, migratory cuckoos that flew into television transmitting towers in the early 1970s had tissues with significant amounts of chlorinated hydrocarbons. Concentrations of pesticide

14 The Cuckoos metabolites in Yellow-billed Cuckoos were higher in autumn than in spring, and were higher in autumn adults than in juveniles, so the time and context suggest that pesticides were acquired on the breeding grounds (Grocki and Johnston 1974). Finally, Greater Roadrunners in North America have disappeared from urban areas and highways as a result of human impact, as well as disturbance and predation from feral and domestic dogs and cats (Emlen 1974). Roadrunners need extensive areas of undisturbed open shrub and grassy habitats with abundant populations of reptile and insect prey, and these are destroyed with the development of agriculture, urban areas, and the fragmentation of expansive tracts of grassland and shrub lands. Few cuckoos are kept in zoos and none have bred in captivity on a large scale that could provide enough birds to reintroduce cuckoos into areas where they have been lost in the wild and where the habitat remains. Cuckoos that have been maintained in captivity where the bright bare colors of their heads and their bright plumage can be well seen by visiting people, include Red-billed Malkoha Zanclostomus javanicus and Chestnut-breasted Malkoha Phaenicophaeus curvirostris. Cuckoos are difficult to breed in captivity

(Pagel 1992), although a few cuckoos, (Great Spotted Cuckoo Clamator glandarius, Common Cuckoo, Guira Cuckoo, Greater Roadrunner and Coral-billed Ground-cuckoo) have bred at zoos, including the National Zoo in Washington, D.C. (Muller 1971) and Vogelpark Walsrode in Germany (Marcodes and Rinke 2000). Captive breeding has been useful for behavioral observations, especially in the Old World ground-cuckoos: behavior and breeding biology are better known through avicultural studies than on the field, where these birds are seldom seen and are nearly unknown (Robiller et al. 1992). Most cuckoos are shy and skulking with humans. The most easily watched in the field are the roadrunners of the southwestern United States and the social group-living Guira Cuckoos and anis in the New World tropics, where these birds are often tame and allow a close approach. Cuckoos are not agricultural pests, and the brood-parasitic cuckoos are not threats to the breeding populations of their hosts. Cuckoos are most interesting to naturalists and biologists who are fascinated with their variety of behavior, the evolution of their sociality, their mating systems and breeding styles which include monogamy, cooperative breeding and brood parasitism.

3 Behavior

Social and parental behavior Cuckoos are solitary, skulking birds, more often heard than seen. Their calls in forests and woodlands are among the most conspicuous and compelling sounds of the tropics. As a group the cuckoos are arboreal and spend their time in trees, although several couas of Madagascar, most coucals in the tropics and some cuckoos in semidesert areas are terrestrial. Many cuckoos live by themselves and call only in the breeding season, while a few kinds of cuckoos live in social groups, leaving their group perhaps only once throughout their lives, when they disperse and join another social group. Cuckoos are mainly diurnal, but both the nesting cuckoos and the brood-parasitic cuckoos call at night along with the nocturnal owls and nightjars. It is in their social and parental behaviors that cuckoos are particularly variable, with monogamous pairs that rear their own young, cooperative breeding birds that rear each others’ young, and brood-parasitic cuckoos that leave their young to be reared in the nests of other species of birds. Many cuckoos rear their own young. Some cuckoos are highly social, living together in permanent social groups where they feed and breed together. Several pairs build a nest and rear their young together in a common nest.These cooperatively breeding cuckoos, the anis and the Guira Cuckoo, live in the New World. The broodparasitic cuckoos, however, have no obvious longterm pair bond. Male and female live apart from each other and come together only to copulate. There is also no family bond between offspring

and their own parents. Females lay their eggs in the nest of another species, the host, and the hosts are foster parents to the young cuckoos. Female cuckoos also remove a host egg to make room for the nestling cuckoo which then removes the others by evicting them from the nest, pushing its own naked body under the eggs and lifting them over the rim of the nest. This behavior ensures that all the parental care of the foster parents will be delivered to the nestling cuckoo.To the host pair the nestling cuckoo means the loss of their own nesting effort. Many cuckoo hosts remove an egg from their nest if it looks unlike their own eggs, and this rejection behavior is the main line of defense against cuckoo parasitism.When the young cuckoo hatches, however, the foster parents accept the cuckoo and rear it at the cost of their own young.This host-parasite association has led to evolutionary battles which each member of the association has won in turn. First the hosts evolve a sensitivity to the cuckoo’s egg in their nest and remove it. Then the cuckoos evolve the ability to match the color and pattern of the host egg (Davies 2000). The nestling cuckoo takes the theme of sibling rivalry to its infanticidal extreme by killing its nestmates. In these broodparasites the rivals, the offspring of the foster parents, are not kin but are competitors for parental care. Cuckoos affect the breeding success of their hosts, both when the female cuckoo removes a host egg from the nest, and when the nestling cuckoo removes the rest of the hosts’ genetic interest by evicting the eggs and nestlings from the nest. Much of our interest in cuckoos and their eggs and behaviors is focused on this brood-parasitic behavior.

16 The Cuckoos

Food and feeding Cuckoos are generalist predators and mainly take large insects. The diet of many cuckoos consists of noxious, brightly colored and hairy forms of caterpillar. These include urticating caterpillars, whose hairs pierce the skin and histamines causing itching and burning, at least when they contact human skin. Cuckoos also take grasshoppers and locusts, many with distasteful “tobacco” juice; other kinds with hard legs and shields. Other cuckoo foods include large non-insect arthropods like millipedes, centipedes, spiders and phalangids, and terrestrial snails and small vertebrates such as tree-frogs.A few cuckoos concentrate on lizards. Roadrunners take many reptiles, and the lizard-cuckoos on Caribbean islands where there are no lizard hawks, feed on lizards. In addition, the brood-parasitic cuckoos take eggs and nestlings from the nests of their hosts, and the coucals take nestling birds when they find them. Old World malkohas and the arboreal New World cuckoos forage by moving through the vines and branches of tropical forests and thickets with hops, twisting the tail for balance like a squirrel, then turning and seizing an insect spotted on a leaf or axil. Most Old World brood-parasitic cuckoos feed in much the same manner, perching and peering in a characteristic posture (Figure 3.1). Cuckoos hunt by ambush as they perch motionless for many minutes and peer about from a perch, then dash or fly to a caterpillar when they sight it on the underside of a leaf or on a tree trunk, grab the insect, and return to their perch to eat it.Their feeding behavior is one of “peer and pounce.” Caterpillars have guts full of indigestible and toxic leaf products, and the cuckoos clean the caterpillars and remove these products. The cuckoo bites off the end and wipes the caterpillar back and forth on a branch until the guts are pressed out, or passes it back and forth through the bill to clean the insides, then swallows the insect. Or else it beats hairy caterpillars against a branch, removing their gut contents and toxins before eating them. These gut removal techniques do not remove the caterpillar hairs. The cuckoo swallows the hairs, which form a felted mat lining in the stomach, and the hairs are regurgitated as a pellet.

Figure 3.1. Fledgling Klaas’s Cuckoo Chrysococcyx klaas, Hans Merensky Nature Reserve,Transvaal.

North American Black-billed and Yellow-billed Cuckoos forage on the tent caterpillars Malacosoma spp. The cuckoos, along with orioles and other birds (60 species are known to feed on these insects, Witter and Kulman 1972), are attracted to wooded areas where the caterpillar tents hang on hawthorns, wild cherries and other rosaceous trees and shrubs.The caterpillars build a communal tent of silk where they retreat between foraging bouts in the tree (Fitzgerald 1995). Although their tent protects the caterpillars, many insectivorous birds pierce the tent and remove the caterpillars. Birds also feed on them as they move along marked foraging tracks. Cuckoos often nest in sites with an abundance of tent caterpillars, M. americanum in spring and M. disstria in autumn. Outbreaks of tent caterpillar in the northern states can defoliate trees each year for up to six years; in the southern states the forests can be defoliated annually for 20 to 30 years (Fitzgerald 1975). The clutch size of Blackbilled Cuckoos varies with the abundance of caterpillars, with large numbers of eggs laid when the

Behavior 17 number of caterpillars is high (Forbush 1927, Sealy 1978). The variation in tent caterpillar abundance may affect the numbers of breeding cuckoos. It is unlikely, however, that cuckoos control the populations of pest caterpillars, although over a few days a cuckoo can take hundreds of caterpillars and clean out a tree of all its caterpillars. Other large insects that appear in outbreaks are also taken, including larvae of gypsy moths Lymantria dyspar ( Jauvin 1996). Caterpillars on economically important trees are more often controlled with the application of chemical pesticides, and bacteria and viruses as biological control agents. The chemical pesticides appear in cuckoo tissues, and from the cuckoo’s point of view it is not a good idea to spray these pesticides in the woods. Coucals take large insects and small vertebrates including snakes, lizards, tree frogs, mice and rats, and small birds.These large cuckoos feed mainly on the ground, often in thick scrub and less commonly in tangles of vines and on the branches of trees. They are generally slow and clumsy in movement but move rapidly when they locate their prey. When hunting grasshoppers and lizards on the ground, they move forward with a slow stalking walk. Then when close, they change to a hop and run, and ambush or chase the prey. Their feeding behavior is one of “flush and rush.” At other times, they tear open bird nests to get the eggs and young, and rip bark from trees to get the lizards and large insects that are concealed underneath the bark. Small birds recognize coucals as predators and they mob the larger coucals. In marshes and swamps, coucals feed partly in the water, where they take frogs, crabs, and aquatic insects, and occasionally they scavenge for dead fish. In seasonally dry habitats, coucals respond to opportunity when fires burn the grass during the dry season. Like foraging Cattle Egrets Egretta ibis, White-browed Coucals Centropus superciliosus approach the smoke and feed at the edge of grass fires, where they take large insects and small mammals that attempt to escape the flames as the fire advances. Other foods include snails, which are rich in protein and in calcium. Senegal Coucals C. senegalensis take large, brightly colored, apparently aposematic bush locusts that give warning displays. They attack and

subdue the locust, then soften it with the bill, working the abdomen and thorax. They then hold the insect underfoot and remove the head and digestive tract and eat the remainder (Goodwin 2001). Anis and Guira Cuckoos are mainly ground feeders. Groove-billed Anis and Guira Cuckoos live in meadows and pastures in dry open habitats, Smooth-billed Anis in denser herbs in wetter habitats, and Greater Anis in wet forest edges. Ani species that live in habitats with larger and wetter leaves have the largest bill: The anis use the bill as a wedge to move through leaves and feed (Willis 1983b), to open the soft wet earth and grab burrowing insects, and to plow through cow dung to get insects (Gosse 1847). The name of the ani genus Crotophaga means tick-eater, a behavior that was more common in the days before chemical pesticides were used in management of livestock.An old German term for anis is “Madenfresser”, eater of ticks (Leverkühn 1894). Groove-billed Anis take grasshoppers and follow cattle much like the icterid cowbirds. Anis also take tuks from the backs of cattle. Anis bite hard-bodied insects repeatedly before they swallow or feed the insects to their young. Smooth-billed Anis sometimes feed on butterflies. In a rapid approach the bird runs to a group of butterflies, lunges, captures one, swallows it, then tries to capture and swallow another, as long as the butterflies remain: Or the bird walks slowly in a crouch towards a single butterfly, then lunges and grabs it. Adult anis are more successful at capturing a large butterfly than are the young birds. Members of an ani group forage near each other, and when they are out of sight of each other in the grass they stay in contact with loud calls. Anis feeding with a cow or horse hop near the front of the mammal and seize insects stirred up in the grass.They catch more insects per minute when they feed with cattle than when they hunt alone, especially in the dry season. Anis in the forest use ant swarms as beaters, catching roaches and other insects as they are driven from the litter by the ants (Sutton 1951, Rand 1953, Smith 1971, Skutch 1983, Willis 1983b, Burger and Gochfeld 2001). Roadrunners feed on the ground. They flush prey as the birds walk with the head low and brush

18 The Cuckoos the grass: The disturbance startles the insects, and when they jump, the roadrunner grabs them. They also flush out insects when they jump up and flap their wings slowly over the ground, then catch the insect in flight or watch it land and pick it from the vegetation.They dart into clumps of vegetation for large insects such as grasshoppers and crickets, spiders as large as tarantulas, and scorpions, and they take vertebrates such as small snakes, mice, small birds and the contents of bird nests.They run down lizards in fast pursuit. Roadrunners leap at flying insects and birds such as swifts that swoop over a dry stream, and they ambush hummingbirds at nectar sites where these visit and feed. They take young bats that fall from the ceilings of caves. In winter they take beetles and grasshoppers, finding insects around the base of rocks.They also take fruits, seeds and scattered grain when insects and reptiles are scarce. Roadrunners eat the sweet seedy fruit of the prickly pear cactus Opuntia, knocking it to the ground and tumbling it to remove the spiny cover: These fruits can make up 10% of a bird’s diet in fruiting season. Following a hand-reared free-living roadrunner through the Texas brush, Sutton (1913, 1915) recorded the food it took in a day: 263 hopping grasshoppers and 73 flying grasshoppers, 17 scorpions, 28 sowbugs, 7 caterpillars, 3 chrysalids, 14 angleworms, 39 moths, 1 butterfly, 14 centipedes, 16 spiders, 2 tarantulas, 3 walking-sticks, 3 small toads, 3 horned lizards, 14 other lizards and a mouse. When roadrunners encounter a rattlesnake, they crouch and circle it, then dart in to grab it behind the head, which they hold in their bill. Then they whip and pound the snake against the ground or stones until it no longer moves. After it is dead they pull the snake into the shade of a shrub and eat it. They also toss and batter other snakes and lizards in this way to kill their prey. Birds are killed and plucked, small mammals are killed by a blow to the base of the skull, and larger prey such as lizards, snakes and ground squirrels are held in the bill, then swung and hit repeatedly against a rock, stick or the ground. The beating kills the prey, separates and crushes the skeleton, compacts the meal and allows the bird to swallow its prey whole. Food processing may take up to 15 minutes. Roadrunners kill scorpions by biting at the tail where the poison gland

and sting are located; this procedure removes the stinger. They swallow lizards and snakes headfirst, avoiding the backwards-pointing spines and scales. They take these reptiles especially in the breeding season when there are young roadrunners to be fed: The pair sometimes cooperates in attacking a snake, circling it and alternating their attacks (Figure 3.2). In the forests of tropical America, groundcuckoos feed at ant swarms, using the ants as beaters to flushout insect prey. Ground-cuckoos are the largest birds in attendance at army-ant swarms and are regular at the swarms along with other forest birds. The cuckoo waits until an ant swarm enters its territory, then follows as the ants pass through the territory and flush out insects with their advance.The cuckoo stands on the ground or perches on a low branch, then runs along the edge of the swarm, snaps up an insect or other forest floor arthropod and runs away, active as it bounds, changes direction or spins: Occasionally it flaps over a vine or gains a higher perch, but extended flight is uncommon. One cuckoo feeds at a time at an ant swarm. When two appear near the swarm the birds either go their separate ways, or they rush at each other with rapid bill-pops, or lower the primaries and spread the tail in aggressive behavior till one bird leaves the site. At other times the ground-cuckoo wanders through the forest, and when it comes to an ant trail it follow the trail until it comes upon the ant swarm. Like a roadrunner, a hand-reared or habituated ground-cuckoo becomes tame enough for an observer to follow on its foraging rounds (Oniki and Willis 1972, Willis and Oniki 1978).

Figure 3.2. Greater Roadrunner pair killing a rattlesnake (after Meinzer 1993).

Behavior 19 A few other cuckoos form a feeding association with ant swarms, or primates and other large mammals, or mixed-species bird flocks. New World ground-cuckoos Neomorphus and Old World ground-cuckoos Carpococcyx follow foraging groups of peccaries and wild pigs, to the extent that these large ground-feeding birds in South America and in Borneo are locally known as “pig birds”. This behavior may have been more common in earlier days when large herds of pigs moved through the forests. Old World coucals sometimes feed in mixedspecies bird flocks. In Madagascar the Blue Coua Coua caerulea follow troops of lemurs as they forage on the forest floor.African yellowbills Ceuthmochares accompany other birds and squirrels, which use each other as beaters.A bird grabs the insects disturbed by the others (Brosset and Erard 1986). Bates (1930: 190) described this behavior: “In its hunt after insects it does not go alone, but follows the squirrels which live in these creepers, or joins itself to other birds, of many kinds, that seek insects, and catches those that flee from the rest; for all insect-catching birds of the trees of the forest find it to their advantage to feed in company.” Yellowbills are seen in 10% of mixed-species flocks in lowland Cameroon (Thomas 1991).Yellow-billed Malkohas Rhamphococcyx calyorhynchus follow Bay Coucals Centropus celebensis through the rattan creepers in Sulawesi, and they also follow troops of primates, Sulawesi macaque Macaca nigra and booted macaque M. ochreata. Redfaced Malkohas Phaenicophaeus pyrrhocephalus sometimes associate with hornbills and other species. Other malkohas that feed in mixed-species flocks include Raffles’s Malkoha Rhinortha chlorophaea. Black-bellied Malkoha P. diardi, Chestnut-bellied Malkoha P. sumatranus and Chestnut-breasted Malkoha P. curvirostris. New World cuculines usually feed alone but some feed in mixed-species flocks of birds. Squirrel Cuckoos Piaya cayana follow army ants in Brazil, and Squirrel Cuckoos and Blackbellied Cuckoos Piaya melanogaster in the Neotropical mainland forests and lizard-cuckoos Coccyzus longirostris in the Caribbean occur in mixed-species flocks, as do wintering Black-billed Cuckoos and Yellow-billed Cuckoos in South America. Greater Anis follow ants and squirrel monkey troops. Other anis live in flocks where they associate with cattle

and take insects when they are stirred up by the cattle, while in less disturbed habitats the anis follow native American mammals. Brood-parasitic cuckoos appear in mixed-species flocks in the non-breeding season. Several insectivorous cuckoos in New Guinea do this, as do wintering Oriental Cuckoos in Australia. Birds in mixed-species flocks benefit from the foraging activities of other birds, and they may also benefit from protection against predators (Thiollay and Jullien 1998). Fruits (especially figs, also berries, tamarinds and palm oil fruits) are the main food of the broodparasitic Dwarf Koel, Common Koel and Channelbilled Cuckoo, and fruit is important in breeding behavior in the Dwarf Koel and Common Koel when the male feeds his mate in courtship. Figs are especially important as they are high in calcium and the fruits are often present in all seasons. Koels have a wide mouth gape (⬎ 2 cm) and can eat as many as 68 intact figs before they regurgitate a pellet that contains the seeds. The young of these broodparasitic cuckoos are often fed insects, although nestling Common Koels reared by Figbirds Sphecotheres viridis get fruit as well as insects, and young Channel-billed Cuckoos Scythrops novaehollandiae are fed fruit by Pied Currawongs Strepera graculina. African Emerald Cuckoo Chrysococcyx cupreus parasitize the fruit-eating Yellow-whiskered Bulbul Andropadus latirostris, as well as insectivorous host species.The bulbuls feed insects to the nestlings during the first four days after their brood has hatched, then feed a mix of insects, regurgitated fruit pulp and whole fruits during the last week of nestling life. Fruit is also fed to young nestling Jacobin Cuckoos when they are reared by fruit-eating bulbuls (Liversidge 1971). Fruit is important to several nesting cuckoos as well. Some malkohas, couas, coucals and Old World ground-cuckoos (Carpococcyx “fruitcuckoos”) take fruit (Shanahan et al. 2001). The ground-cuckoos in Sumatra and Borneo live in dipterocarp forests which produce huge crops of fruit in some years, but in most years do not produce any fruit ( Janzen 1974, Appanah 1985). The fruiteating animals such as bearded pigs Sus barbatus that take these masting fruits have to move long distances to find food or switch to alternative foods such as figs (Whitten et al. 1987, MacKinnon et al. 1996).

20 The Cuckoos Ground-cuckoos often feed with large mammals nearby. Dipterocarp forests are heavily logged and burned and during El Niño years the trees fail to fruit (Curran et al. 1999, Curran and Leighton 2000). Dependence on the fruit of these masting trees may be responsible in part for the decline of groundcuckoos in the Greater Sundas. In Madagascar, two couas are vegetarians that take mainly fruit; couas also take the gum that exudes from acacia trees. In the New World, roadrunners, Squirrel Cuckoos and Yellow-billed Cuckoos occasionally take fruit along with a mainly insectivorous diet. Koels and other fruit-eating birds are ecologically important in the dispersal of forest fruits (Lambert 1989, 1991, Lambert and Marshall 1991, Compton et al. 1996, Kinnaird et al. 1999, Shanahan et al. 2001). Smooth-billed Anis take hot peppers Capsicum, the “Madame Jeanette” chili peppers of the Caribbean and Suriname.The capsaicin chemical in the ripe fleshy fruit is distasteful to mammals but not to birds. The function of capsaicin is to deter mammal seed predators (small rodents are fruit consumers which crush the seeds and do not allow them to germinate) and not to deter birds, which are attracted to the bright fleshy fruits and disperse the seeds in their feces. Consumption of non-pungent peppers by fruit-eating mammals results in zero germination. In contrast, consumption of both non-pungent and hot peppers by fruiteating birds results in germination rates similar to those of control seeds planted from the fruit, which proves that consumption by birds results in successful dispersal of the seeds (Tewksbury and Nabham 2001).Anis disperse gumbo-limbo fruits when they remove the fruit skin, swallow the flesh and regurgitate the seeds.

Cuckoos and their predators Larger predators such as raptors and large primates take adult cuckoos, and cuckoo nestlings and eggs are taken by a variety of predators including birds, mammals and snakes. Both young and adult broodparasitic cuckoos are eaten by mammals such as monkeys in Africa.Young roadrunners are eaten by dogs, cats and raccoons, and these mammals also catch adult roadrunners on the nest. Black-billed

Cuckoos in North America are eaten by hawks and other predatory birds (Hughes 2001), Cuculus cuckoos in the Philippines are plucked and eaten by raptors (Curio et al. 2001), anis in Central America are taken by carnivorous bats (Vehrencamp et al. 1977), and coucals in Central Africa are eaten by chimpanzees Pan troglodytes (Nishida and Uehara 1983), all observations suggesting that cuckoos are not toxic to predatory birds and mammals. In addition, rural peoples eat cuckoos for medicinal purposes as well as for food, including anis in the Caribbean, Common Koels in India and coucals in New Guinea (Walden 1869, Payne 1997b). Most cuckoos avoid predators by stealth. The arboreal New World Black-billed and Yellow-billed Cuckoos avoid predators by their watchful behavior and are easily overlooked by predators, as well as by their visually-oriented prey. Other cuckoos actively avoid predators.When they see a predator, group-living anis call and fly together to a higher perch, so their group behavior is related not only to food but also to mutual protection from predators. Adult Greater Roadrunners avoid an aerial predator by running, dodging, flashing their wings and spreading their tails, behaviors that may distract, confuse or intimidate a predator (Rand 1941b). Roadrunners often run rather than fly when a predator approaches. When I taught a field ornithology course with field trips in the Black Mesa country of western Oklahoma, we watched where the bird ran—it hid in a clump of cholla cactus.The students circled the cactus, one reached in and grabbed the roadrunner. Then we took a few photos, released the bird and watched it run again. Nestlings of many nesting cuckoos (Black-billed and Yellow-billed Cuckoos, roadrunners, coucals) void a foul black liquid from the cloaca when they are disturbed in the nest, and this may repel a predator. Nestling Guira Cuckoos of any age sometimes leap from the nest or branches and escape a predator when they land safely on a lower branch and hide; leapers sometimes die from the fall (R. Macedo, in litt.).The young cuckoos are at risk: more than half of all roadrunner nests are lost to predators (Folse 1974).Young nestling roadrunners produce a fecal sac that is removed and eaten by the parent cuckoos; older nestlings when about to

Behavior 21 leave the nest produce the foul liquid when a predator disturbs them. Brood-parasitic young cuckoos do not produce the black liquid, but like their passerine host young they encapsulate their feces and urine in a gelatinous fecal sac which is removed by the foster parent.Although nesting cuckoos take an active role in mobbing predators, the broodparasitic cuckoos do not join these mobs; they leave the defense of the young to the foster parents.

Breeding seasons In temperate latitudes the gonads of many kinds of birds develop in response to the photoperiod. Seasonal daylength changes are reliable predictors of environmental conditions such as food and temperature, which affect the ability of birds to rear their young (Murton and Westwood 1977). In the northern temperate New World, brood-parasitic Brownheaded Cowbirds Molothrus ater breed at the same time as their host species. In experiments, the cowbirds respond to seasonal changes in daylength in the same way as their hosts, and in the same way as nesting icterids in the same region (Payne 1969b, 1973b). It is likely that brood-parasitic cuckoos respond to the same environmental cues as their hosts in the initial stages of hormonal activity and gonad growth. The final stages of ovarian growth of the brood-parasitic birds depends on hearing the songs and seeing the nesting activity of their hosts (Immelmann and Immelmann 1967, Payne 1973b, 1983, Löhrl 1979). No experimental work on this has been carried out with cuckoos, but it is likely that they respond to changing conditions in much the same way as these other brood-parasitic birds do. North American Black-billed and Yellow-billed Cuckoos breed along with the earliest songbirds in April and May, and if insects are still abundant they continue to breed into August and early September. These early, long and late seasons occur notably in areas with outbreaks of tent caterpillars, which feed in huge numbers on many wild rosaceous trees and shrubs. For Greater Roadrunners, breeding varies with temperature and rainfall. In southern California they breed from late February in low elevations of the Lower Sonoran Life Zone,

and a month later in the upland deserts of the Upper Sonoran Life Zone. A pair can nest again when the male takes over the care of the fledged brood while the female lays a second set of eggs.The cooperative behavior of a pair may let them rear more than one brood when insect food is abundant during the short desert season. In southern Arizona, breeding tends to be bimodal with nesting from mid-April to mid-June and late July to midSeptember with a pause in the hot dry summer, and nesting again after the summer rains, with the breeding peaks varying with the rains of the year. In Texas they breed from March to October, and in Oklahoma from April to July or August. In the tropics, cuckoos breed in the rains; many cuckoos in anglophone regions of the world are known as “rain birds”. They begin to sing shortly before the first good rains begin, and their calls are taken as cues to local farmers to prepare their land for cultivation. Birds may begin their physiological change into breeding condition well before the rains begin (Immelmann 1971). Although their breeding seasons are indirectly timed with the season of rains, some tropical birds respond directly to changing daylength much as in northern temperate regions. A seasonal change of less than an hour, that is, the daylength change at 15° from the equator, is enough to affect the cycles (Hau et al. 1998). Outside the forest where the rains fall through fewer than six months, breeding is seasonal with changes in the vegetation and insects (Levings and Windsor 1982, Wright and Cornejo 1990). In tropical forests many birds are seasonal, nesting in all but the rainiest or driest season (Chapin 1939, Moreau 1950, Skutch 1950, Fogden 1972, Bell 1982a,b,Wikelski et al. 2000). The breeding season for brood-parasitic cuckoos corresponds with the breeding season of their host species. In tropical and subtropical regions, the cuckoos breed early in the rains when their host species begin to nest.These seasons vary within Africa. (1) In equatorial East Africa the breeding season corresponds to the local rains, where seasons differ by a few months in areas only 200 km apart (Brown and Brown 1980). These differences are associated with altitude, distance from the coast, and the presence of mountains that intercept monsoon rains from the Atlantic. East of the eastern rift (region D)

22 The Cuckoos there are two rains in a year, with long rains in April and May and shorter rains around November: January to March is a hot dry season and June to October is cooler and overcast.West of the rift and south of Lake Victoria (region C) there is only one season with rains between October and April or May, and a dry season in the middle of the year, although temperatures are cooler than in the dry season of D owing to the elevation and to the sun not being overhead at that time. In western Kenya, the Lake Victoria basin and Uganda (region B), rains come from March to November with two peaks corresponding with those east of the rift. In the dry region of northern Uganda and northern Kenya (region A) the rains come between March and October, with little rain in June and July. In coastal Kenya the rains come from April to June during the southeast monsoon. Insectivorous songbirds breed during the rains in these areas, and their brood-parasitic cuckoos breed in the same season as their hosts. Klaas’s Cuckoo Chrysococcyx klaas breed mainly in the rains as do their sunbird hosts, with most recorded from February to April in region D, December to April in region C, April in region B and September in region A. Diederik Cuckoo C. caprius breed for a longer season from March to June when most ploceids breed in region D. Finally, Red-chested Cuckoos Cuculus solitarius breed mainly from April to June in region D, and February to September in region B. (2) In West Africa, seasonal rains from the Atlantic are controlled by the inter-tropical convergence (ITCZ), which is the atmosphere’s equator.The zone moves a month or two behind the overhead movement of the sun.The vernal equinox in March is followed by rains in West Africa in May and these rains continue until October. The zone forms by movement into, or convergence of, the northeast trade winds of the northern hemisphere and the southeast tradewinds of the southern hemisphere in the equatorial belt.The inflowing air rises and clouds form in convective cells that often drift westward bringing frequent rains and thunderstorms. Dry northeasterly winds, the harmattan, blow off the Sahara desert from November to March, signaling the end of the rains in the north.The wet westerly or southwestern monsoon is more extensive and marked in summer

from May to September when thermal heating over the continent supports the northerly displacement of the equatorial trough. Rains on the coast of Ghana at Accra and Cape Coast peak in May and June, as the monsoon is intercepted along the coast west of Cape Coast. Further east in Togo the rains are affected by a low north-south mountain range (Grimes 1987, Cheke and Walsh 1996). Only 200 km inland from the coast in Ghana at Kumasi are there two rainy peaks, one in May and June and the other in September and October (Elgood et al. 1994). Further north the winds bring less rain. In coastal Togo, Lomé gets nearly 100 mm during April, while in northern Togo rains this intense arrive a month later and the growing season begins later (Cheke and Walsh 1996). Likewise northern Senegal gets less rain and the season of effective rain is from June to October (Morel and Morel 1990). In Nigeria a monthly rainfall of at least 100 mm comes a month earlier on the coast (March) than at Ibadan (April), and a month later in northern Nigeria (May) (Elgood 1994). Different breeding seasons are known for African Black Coucal Centropus grillii—in coastal Ghana from April to July, in northern Ghana in July and August.The birds also apparently migrate between north and south (Grimes 1987), much as they do in Togo. (3) Even at high latitudes the breeding seasons are strongly influenced by seasonal rains. Klaas’s Cuckoos in the Cape Region of South Africa, an area of winter rainfall at 34–35° N, breed in winter from July to September along with their sunbird hosts, and in the Eastern Cape at 33–34° N they breed mainly from October to December. Further north (c. 15° S) where rains come later the cuckoos breed in the later months, as in Zambia where they breed mainly from December to February. Great Spotted Cuckoos breed two or three months later in western Namibia, than in Natal and the Eastern Cape of South Africa, again in relation to the rains of the region. In India the breeding seasons of brood-parasitic cuckoos are similar to those of nesting cuckoos. Their breeding seasons are related to altitude in India (Baker 1934, 1942), where cuckoos and their hosts breed earlier in the plains than in the higher foothills and mountains.The rains and green growth that follow come earlier at the lower altitudes.

Behavior 23 In Africa the breeding season for the broodparasitic cuckoos varies with the ecology of the nesting cuckoos as well as with the ecology of the host species. In Malawi, nesting resident Whitebrowed Coucals breed right through the year except from July to September, while migratory African Black Coucals breed from January to May when the marshes are flooded (Benson and Benson 1977). Many insect-eating songbirds begin to breed before the first rains, and in northern Zambia like the miombo trees in Brachystegia woodland which grow and unfold their new leaves before the rains, these songbirds anticipate the rains weeks before they begin. In this pre-rain period, Black Cuckoos Cuculus clamosus and Klaas’s Cuckoos begin to sing. The nesting cuckoos tend to have a longer breeding season than the brood parasites, depending on their status as migrants or as residents (Table 3.1). In the parasite species with the largest number of records, the season is five months. In Zambia, the cuckoos with the long breeding season are resident coucals. In Zimbabwe, the brood-parasitic cuckoos with the longest season are Levaillant’s Cuckoos Clamator levaillantii, whose main host,

Table 3.1 Breeding seasons (number of months) in south-central Africa.

nesting cuckoos Centropus grillii Centropus cupreicaudus Centropus senegalensis Centropus superciliosus Ceuthmochares australis brood parasites Clamator jacobinus Clamator levaillantii Pachycoccyx audeberti Cuculus clamosus Cuculus solitarius Cuculus gularis Chrysococcyx caprius Chrysococcyx klaas Chrysococcyx cupreus

Malawi

Zambia

Zimbabwe

5 na 1 9 1

3 6 7 3 -

4 4 6 6 1

7 2 1 3 1 7 5 3

5 3 4 3 5 5 -

6 8 5 4 4 4 6 8 -

Arrow-marked Babbler Turdoides jardinei, is a cooperatively-breeding bird that nests in all months; and the partly resident Klaas’s Cuckoos whose hosts are warblers and sunbirds, some breeding in the rains and others in the dry season. The resident coucals breed over a longer period than the migratory African Black Coucals (Irwin 1981, Hustler 1997a). In Australia, the time when cuckoos breed depends on the rains. In two regions matched for latitude but with different months of rain, Horsfield’s Bronze-cuckoo Chrysococcyx basalis begins to lay in the early austral spring in July in the southeast, and in early October in the southwest. A common host in the west, Splendid Fairy-wren Malurus splendens have a longer breeding season than their cuckoos, and the latest nests escape the cuckoo parasite (Rowley et al. 1991, Rowley and Russell 1997). Further north where rains fall in austral summer, cuckoos and their hosts breed in the Kimberley Division from January to March, and from March to May and August to September in the Pilbara Region: In arid inland Australia breeding is bimodal, with rains in spring and autumn (Storr 1984a, Brooker and Brooker 1989a,b, Ambrose and Murphy 1994, Pizzey and Knight 1998, Higgins 1999).

Migration Cuckoos undertake long-distance migrations between their temperate breeding areas and their wintering grounds.With cuckoos that breed in the higher latitudes, migration is known from their alternating seasonal occurrence in the breeding region and on the winter grounds. North American Black-billed and Yellow-billed Cuckoos migrate to South America, and winter south of the equator. Another North American cuckoo is at least partly migratory: northern populations of Mangrove Cuckoos appear seasonally in Florida and the Bahamas, and are absent in winter. Migration is just as remarkable although the detail is less well known with the South American cuckoos. Austral migration occurs with other birds that breed in temperate South America while the cuckoos are nonbreeding wintering birds in the tropics (Chesser 1994, 1997). At least three cuckoos have

24 The Cuckoos this seasonal movement and tend to live in scrub woodland both in their breeding grounds and in the austral winter: Ash-colored Cuckoo and southern populations of Pearly-breasted Cuckoos Coccyzus euleri and Dark-billed Cuckoos C. melacoryphus. Ash-colored Cuckoos Coccycua cinerea breed only in the southern temperate region, but the other two cuckoos have a wide breeding range. Other cuckoos perhaps migrate from the southern extremity of their range (Greater Ani, American Striped Cuckoo: Chesser 1994), yet the odd birds that appear in the tropics in localities where they are not regular in occurrence may be local dispersers rather than austral migrants. In the absence of known morphological differences between northern and southern breeders it has not been possible to determine whether the southern cuckoos move into the northern region while northern birds are breeding, or whether both populations spend a non-breeding period in the Amazon basin. Four cuckoos in the northern temperate region of the Old World complete long-distance migrations when they move across the equator (Great Spotted Cuckoo, Common, Oriental and Asian Lesser Cuckoos). European Great Spotted Cuckoos and Common Cuckoos migrate to Africa, flying nonstop over the Mediterranean and the Sahara, a distance of more than 3000 km. Other cuckoos migrate from continental Asia to the Malay Peninsula, the Greater Sunda Islands and the Philippines (hawk-cuckoos Hierococcyx and drongocuckoos Surniculus), or over the Indian Ocean to Africa (Indian populations of Jacobin, Common and Asian Lesser Cuckoo) and appear in migration on islands in the Indian Ocean between their breeding and wintering grounds. Some cuckoos that appear on the coast in Kenya have flown over 3000 km from their nearest breeding grounds to winter in Africa. Madagascar Lesser Cuckoos Cuculus rochii migrate from their breeding grounds to live the rest of the year in continental Africa. Migration of the Common Cuckoo is known both by seasonal changes in where the birds live, and by recoveries (n ⫽ 2) of birds marked in Europe and found thousands of kilometers away in Africa in the winter. Less well known are the wintering grounds of the Common Cuckoo that breeds in eastern Asia.

Migration of brood-parasitic cuckoos that move between breeding and nonbreeding areas is timed to coincide with the season when the hosts breed, and when caterpillars emerge and become abundant with the onset of warmer weather in spring in the high latitudes, and with rains in the subtropics and tropics. Cuckoos regularly migrate within Africa, moving from their breeding grounds in South Africa to equatorial Africa (black phase Jacobin Cuckoo, Black Cuckoo). Cuckoos move with the rains in West Africa where they normally live near the humid coast in the dry season, and move northward into the savanna in the rains ( Jacobin Cuckoo, Levaillant’s Cuckoo, Black Cuckoo, African Emerald Cuckoo, Klaas’s Cuckoo, Diederik Cuckoo, African Black Coucal), while some may remain as residents of coastal areas throughout the year. In the Australo-Pacific region, certain cuckoos migrate from their southern breeding grounds in the temperate regions of Australia and New Zealand, to winter in the tropics (Horsfield’s Bronze-cuckoo, Shining Bronze-cuckoo, Channelbilled Cuckoo, Long-tailed Cuckoo). Migrating cuckoos are often very fat, and their stores of fat may persist into winter (“summer” in their nonbreeding grounds), as in a Common Cuckoo from the Palearctic measured in December in South Africa (Payne 1968, Moreau 1972). Shining Bronzecuckoo and Long-tailed Cuckoo breed in New Zealand in summer but are absent there in the winter months; and they occur on tropical Pacific islands in winter but not in summer. Few banded cuckoos in the region have been recovered to show longdistance migration, In one case, Common Koel banded in New South Wales, Australia, was recovered in New Guinea, 2950 km from its banding site (Higgins 1999). The only form of Little Bronzecuckoo known to migrate from Australia to New Guinea is Chrysococcyx m. barnardi, with a few winter records in New Guinea. Other Australian bronzecuckoos move within Australia but it is unknown whether the green-glossed C. m. minutillus migrate to New Guinea. Rufous birds that breed in Australia look like rufous breeding birds in New Guinea (C. m. poecilurus), so there is no compelling evidence of long-distance migration in these cuckoos.

Behavior 25 Migration of cuckoos that breed in tropical areas is less well known, and cuckoos seen only in one season, may be resident but inconspicuous at other times when they are silent. In Africa, Jacobin Cuckoos and Levaillant’s Cuckoos are seen seasonally in Zambia within 15° of the equator and they migrate to more equatorial regions. Klaas’s Cuckoo is conspicuous in the breeding season in southern Africa and nearly disappears in the non-breeding season. Nevertheless a few birds remain all year, but are inconspicuous during the season when they do not sing. Movements of cuckoos within Africa are not well understood, partly because several species have populations that do not differ morphologically between regions and lack distinctive subspecies, and in these birds there is no easy way to distinguish between local residents and nonbreeding visitors. The black-phase Black Cuckoos in West Africa probably come from populations further south (that is, south of the Congolese equatorial forests), if only because these black-phase birds are not known to sing or to breed in West Africa. On the other hand, there are remarkably few records of the resident forms of this species actually breeding in West Africa.These migratory forms are mainly non-forest birds. In Togo, five species of non-passerines in forest habitats show seasonal changes in their numbers, that suggest a seasonal arrival of birds from further south: three of these are cuckoos, Black Cuckoo, Red-chested Cuckoo and African Emerald Cuckoo. Among savanna cuckoos several species appear to have resident populations in the south, but these birds are present only during the rains in the northern savannas and they may undertake more local migrations. These include Diederik Cuckoo, Klaas’s Cuckoo and African Black Coucal, and the coastal swamp grasslands may also receive visiting African Black Coucals from further south (Cheke and Walsh 1996). In East Africa, Barred Long-tailed Cuckoos Cercococcyx montanus appear seasonally near the east coast where they do not breed, and these birds make local migrations. In southeast Asia, the migration of cuckoos at night has been recorded in the Malay Peninsula. On Fraser’s Hill, the records of cuckoos netted at night include more than 10 individuals each of Chestnut-

winged Cuckoo Clamator coromandus,Violet Cuckoo Chrysococcyx xanthorhynchus, (Fork-tailed) Drongocuckoo Surniculus (dicruroides) (the most abundant, with 57 records),Whistling Hawk-cuckoo Hierococcyx nisicolor and Indian Cuckoo Cuculus micropterus (Medway and Wells 1976). The seasonal appearance of small glossy-cuckoos Chrysococcyx in the Indonesian archipelago may result from the inconspicuous behavior of non-breeding birds, and there are no substantiated records of migration of the distinctive forms of Little Bronze-cuckoo C. minutillus. Brush Cuckoos Cacomantis variolosus may migrate from the central highlands to the lowlands in New Guinea: the evidence is the occasional appearance near the coast of long-billed birds, that are more often found in the mountains (Diamond 1972). Long-distance migrants sometimes appear on the wrong side of the ocean. North American Black-billed and Yellow-billed Cuckoos appear as vagrants on the North Atlantic islands and in northern Europe, and as far east as Italy and Sicily, while Palearctic Common Cuckoos turn up in northern North America in Alaska and on the Atlantic coast, and Oriental Cuckoos appear in Alaska.The dangers of migration over the ocean are illustrated by the remains of a Yellow-billed Cuckoo in the stomach of a shark (Saunders 1962). Black-billed and Yellow-billed Cuckoos migrate at night.They make an over-water flight in autumn of 2000–3000 km from their breeding grounds in North America to the West Indies, then continue onwards to South America. Or they fly as far as 4000 km from the north directly to the mainland of South America, then continue well inland to winter south of the equator. Yellow-billed Cuckoos appear through May to August on islands off the Florida coast where they do not occur as breeding birds, and their appearance suggests that spring migration continues as late as June, and the southbound return begins by July (Stevenson and Anderson 1994).These cuckoos orient by patterns of the stars: when tested in a planetarium (Marler and Hamilton 1966) they became active at night and directed their movements in the seasonally correct direction of the planetarium sky. These cuckoos appear as casualties below TV transmitting towers which attract nocturnal migrants by their lights,

26 The Cuckoos and they fly into brightly lit walls as well (Sick 1993). Cuckoos that breed in New Zealand, the Shining Bronze-cuckoo and Long-tailed Cuckoo, are nocturnal migrants that winter on remote islands in the Pacific Ocean more than a thousand kilometers from their breeding range. The large Channel-billed Cuckoos migrate during the daytime as they pass between Australia and the tropics where they winter in New Guinea and the Moluccas. Cuckoos like other nocturnal migrants can orient even without the experience of learning from their parents. When they are socially isolated and hand-reared apart from the adults, young Yellowbilled Cuckoos become active at night in the migration season and orient their restless behavior toward the stars (Marler and Hamilton 1966). A “natural experiment” demonstrated that the young brood-parasitic cuckoos have no opportunity to learn the migration time or route from the adults, as the older birds do not live with their offspring, and would have left from their northern breeding

grounds several weeks before the juveniles leave the area. Some resident cuckoos disperse after the breeding season and range far from their natal and breeding areas. Individual Greater Anis appear outside their breeding range in Brazil away from their breeding sites. Smooth-billed Anis and Groovebilled Anis occasionally show up at the end of summer in North America more than 1000 km north of their breeding grounds, usually alone and not in a group. A South American Guira Cuckoo once appeared in the Netherlands Antilles. “All three [ani] species have shown a remarkable tendency towards extralimital vagrancy, given their seemingly ineffective flight and short rounded wings. The wide distribution of Smooth-billed [Ani] in the West Indies with no phenotypic differentiation attests to its powers of dispersal” ( J.V. Remsen, in McLean 1995). Most tropical cuckoos are resident, including the ground-cuckoos of the Old World and New World, the Madagascar couas, the malkohas and many coucals.

4 Morphology

Cuckoos have a zygodactyl foot with two toes of each foot (numbers 1—the inner toe or hallux of most birds—and 4, the outermost toe) directed backward, and the other two toes (numbers 2 and 3) directed forward. Most cuckoos have a long tail. Cuckoos have 10 primaries and 10 rectrices (eight in the anis); the plumage has no aftershaft; the oil gland is naked, and in most species the nestlings have no downy feathers (Nitzsch 1867). Several other morphological features occur in cuckoos that are generally not present in other birds, and a few of these features occur only in the cuckoos.The most distinctive charactereristics are in the skeleton in details of the tarsometatarsal bone (two enclosed canals side by side in the hypotarsus, and a characteristic shape of the accessory process or sehnenhalter on trochlea IV of the distal tarsometatarsus), and in the shape of the humerus and its deltoid crest. Relationships within the cuckoos have been estimated from their anatomy, geographic distribution and breeding biology, and more recently from their molecular genetics. Earlier studies were based on morphology, primarily from museum skins. Shelley (1891) recognized one family of cuckoos with six subfamilies: typical cuckoos Cuculinae, coucals Centropodinae, malkohas Phaenicophaeinae, New World ground cuckoos Neomorphinae, New World brood parasites (the group: the birds were not yet known to be brood parasites) Diplopterinae, and anis Crotophaginae. Peters (1940) recognized the cuckoo family, also with six subfamilies (Cuculinae, Phaenicophaeinae, Crotophaginae, Neomorphinae, Couinae and Centropodinae); he included Carpococcyx with Neomorphus in Neomorphinae. Wolters (1975–1982) recognized nine families of cuckoos

(Crotophagidae, Centropodidae, Neomorphidae (including Dromococcyx), Taperidae, Coccyzidae, Clamatoridae (including Pachycoccyx), Cuculidae, Eudynamidae (including Caliechthrus, Urodynamis, Microdynamis, Rhamphomantis and Scythrops), and Phoenicophaeidae. He included Coua and Carpococcyx with Centropus in a family Centropodidae; Mason (1997a) listed characters that these groundliving birds had in common.

Comparative and adaptive features of cuckoo morphology A few general trends in morphology are related to the life style of cuckoos. In some cases we can distinguish between adaptations and features that are explained by the evolutionary relationships among lineages. Body weight was taken as the main standard of body size to allow broad comparisons across cuckoos: body weight is correlated with other aspects of size, and many life history variables are associated with body weight in birds in general, as in a sampler of behavior, breeding biology, lifespan and metabolic rates (Calder 1984). Body size was then compared with brain size in cuckoos with different breeding behavior (brood parasites, nesting pairs and cooperatively-breeding group-living birds), and also with length of the leg bones in relation to the arboreal and terrestrial feeding behavior of the cuckoos.Where body weight is not known, wing length or even tarsus length can be used as an estimate of body size, at least within a group of birds based on the same body plan such as the coucals, but where the proportions of birds differ

28 The Cuckoos among groups, body weight is the standard of comparison for body size. Species differences in brain size may be related to both phylogeny and adaptive behavioral ecology. There have been few comparative studies of birds (Pearson 1972, Mlíkovsky 1989), and those have shown few ecological correlates of brain size, except that altricial birds tend to have larger brains than precocial birds (Bennett and Harvey 1985, Iwaniuk and Nelson 2001, 2002). In cuckoos, the brain size is similar to that of other altricial birds of the same body size, although the brain is smaller than in parrots.The relationship of brain size with body size in birds varies with the systematic level of analysis, a so-called “taxon-level effect” ( Jerison 1973, Aboitiz 1996, Nealen and Ricklefs 2001). To test the variation in brain size with behavior, skeletal specimens of 76 cuckoo species were examined by George Kulesza. The orbital foramina were closed with masking tape, then the braincase was

filled through the foramen magnum. Spherical lead shots of uniform size were introduced into the braincase with the help of a flexible cylinder.The volume of shot in the braincase was compressed by shaking and tapping the skull until the shot was level with the foramen magnum. The procedure was repeated for each skull to determine repeatability (within 2%) and the two measurements were averaged. The amount of shot was determined with the use of a graduated pipette. Volumetric measurements showed that 0.17 ml of shot was equal to 1 ml or about 1 g of brain. The mean mass of a pellet (0.013 g) agreed closely with the mass µ calculated for a sphere of the measured diameter 0.136 cm (µ ⫽ ρ (4/3)πr3), where radius r ⫽ 0.068 cm, and density of lead ρ ⫽ 10.5 g (cm3)⫺1.Where several skeletal specimens of a species were available, 2–4 skulls were measured and the mean brain size was calculated. Brain size (g) was then compared with body size (g) in the corresponding sex and subspecies.

Figure 4.1. Brain size in relation to body size and behavioral ecology in cuckoos. Closed symbols indicate nesting species (N), open symbols indicate brood parasites (P).

Table 4.1 Leg length, femur, tibiotarsus and tarsometatarsus (mm), brain (g), body weight (g) and behavior (a ⫽ arboreal, t ⫽ terrestrial). Species

breeding (c, n, p)2

specimen3

41.0 43.6 38.3 36.3

t t t t

c c c c

UMMZ UMMZ UMMZ UMMZ

49.8 86.5 70.8 95.1 40.2 46.6 40.5

38.0 63.7 51.2 74.6 30.5 36.2 32.3

t t t t t t t

n n n n p p p

UMMZ 159111; 133738, 159111 UMMZ 227057; 227059, 234551 UMMZ 159112; 156461, 159112 UMMZ 200592 UMMZ 222217; 218370, 218946 MVZ 85638 LSU 101257

37.9 48.4 60.6 52.3 63.5 73.8 54.4 48.8 60.5 41.7 45.1 39.9

55.1 71.7 87.9 79.9 95.3 112.3 74.9 67.5 90.6 59.3 63.5 62.3

41.4 45.6 57.5 50.1 63.4 71.6 51.1 48.4 63.5 41.2 43.1 43.8

t t,a t t t t t t t t t t

n n n n n n n n n n n n

UMMZ 143041 USNM 226188 USNM 557171 USNM 318594 AMNH 7333 UWBM 63149 UMMZ 224761 UMMZ 207709 BMNH S/1969.1.462, USNM 562052 UMMZ 208305 UMMZ 208402 UMMZ 228027; 228025, ⫺26, ⫺27 (s)

64.4 66.3 42.4

112.3 112.4 74.4

87.4 91.7 57.9

t t t

n n n

UMMZ 219851, 236029; ⫺26 and 219043 (s) USNM 223970 USNM 432195; FMNH 356640 (s)

brain (g)

femur

tibiotarsotarsus metatarsus

142 153 105 75

1.85 2.69 1.71 1.53

37.2 41.9 38.3 34.8

58.8 64.6 56.8 53.6

62 305 178 345 52 85 48

1.33 3.59 2.77 4.39 1.25 1.11 0.86

31.4 54.5 50.0 60.8 27.9 30.3 24.3

88 – – 336 503 769 280 170 417 170 177 123

2.25 4.38 5.33 – 7.45 9.69 3.90 3.34 4.66 3.11 3.36 2.75

– – 190

5.96 – 2.63

202015; 157046, 200661, 202013, ⫺14 (s) 139988; 139989, 219553 (s) 218940; 118151, 218943 (s) 219189; 133737, 219189, 219557 (s) (s) (s) (s) (s)

Morphology 29

Crotophaginae Guira guira Crotophaga major Crotophaga ani Crotophaga sulcirostris Neomorphinae Morococcyx erythropygus Geococcyx californianus Geococcyx velox Neomorphus geoffroyi Tapera naevia Dromococcyx phasianellus Dromococcyx pavoninus Centropodinae Centropus bengalensis Centropus celebensis Centropus goliath Centropus leucogaster Centropus menbeki Centropus milo Centropus phasianinus Centropus senegalensis Centropus sinensis bubutus Centropus superciliosus Centropus toulou Centropus viridis Couinae Carpococcyx renauldi Carpococcyx radiatus Coua ruficeps

behavior (a,t)1

body (g)

Species Coua caerulea Coua cristata Coua reynaudii Coua gigas Coua serriana Cuculinae Rhinortha chlorophaea Ceuthmochares australis Clamator coromandus Clamator glandarius Clamator jacobinus Clamator levaillantii Dasylophus superciliosus Dasylophus cumingi Zanclostomas javanicus Phaenicophaeus diardi Phaenicophaeus tristis Phaenicophaeus curvirostris Rhamphococcyx calyorhynchus Coccycua minuta Piaya cayana Piaya melanogaster Coccyzus americanus Coccyzus minor Coccyzus melacoryphus Coccyzus erythropthalmus Coccyzus lansbergi Coccyzus merlini Coccyzus longirostris Coccyzus pluvialis Coccyzus rufigularis

body (g)

brain (g)

femur

235 136 140 420 298

3.48 2.38 3.47 4.75 4.15

50.0 34.7 43.5 57.8 52.5

tibiotarsotarsus metatarsus 83.8 55.4 55.5 39.3 66.4 46.3 98.1 72.0 80.1 61.0

59 66 77 124 72 124 118 174 98 61 115 189 – 40 98 102 63 64 50 51 55 154 98 176 130

1.34 1.53 1.49 1.85 1.31 1.89 2.62 3.27 2.46 – 2.39 2.98 – 1.06 2.07 1.89 1.20 1.53 1.15 0.99 1.12 2.50 1.96 2.79 2.62

28.1 31.0 32.1 35.6 31.7 36.9 40.1 46.1 37.6 34.4 40.6 45.8 49.6 25.6 42.0 38.2 29.2 29.9 27.9 26.2 28.6 42.3 37.4 47.6 48.2

41.3 44.2 45.4 49.6 44.5 51.2 57.3 67.2 51.2 47.2 59.5 65.1 67.1 36.1 61.4 53.7 39.7 43.5 39.0 37.1 41.5 61.6 51.9 68.8 –

27.1 29.9 27.8 32.5 29.0 32.5 38.7 44.0 35.6 33.5 40.5 43.2 45.7 25.4 42.4 37.4 27.4 29.7 26.3 24.6 27.0 41.9 35.1 45.1 –

specimen3

behavior (a,t)1 a a t t t

breeding (c, n, p)2 n n n n n

UMMZ 209201 UMMZ 157526; 157526, FMNH 384680 (s) UMMZ 219043; 208403, FMNH 352797 (s) FMNH 360040 UMMZ 209202; FMNH 356638, ⫺39 (s)

a a a a a a a a a a a a a a a a a a a a a a a a a

n n p p p p p n n n n n n n n n n n n n n n n n n

MCZ 342734 UMMZ 209209; 158185–86 (s) USNM 343999 UMMZ 212911; ⫺11, ⫺14 (s) UMMZ 212899; 212890, ⫺94, ⫺96 (s) UMMZ 212906 UMMZ 228068; ⫺22, ⫺23, ⫺24 (s) UMMZ 233062; ⫺63 (s) UMMZ 236439 BMNH S/1969.1.10 USNM 344002 UMMZ 236028 USNM 226196 UMMZ 139991; ⫺90, ⫺92 (s) UMMZ 222216; 133732, ⫺33 (s) UMMZ 209478; LSU 64990, 89701 (s) UMMZ 203320; 154465, 201937, 207641 (s) UMMZ 208492; 152623, 208492 (s) UMMZ 208493 UMMZ 205253; 200574, 201813 (s) LSU 114314 UMMZ 158527 USNM 554614 USNM 559182 USNM 226541

30 The Cuckoos

Table 4.1 contd.

1

40.9 36.7 58.7 19.6 16.5 16.8 17.2 20.6 18.3 23.0 25.8 20.6 19.2 19.0 22.1 19.1 33.7 29.1

57.1 51.9 80.2 29.7 25.5 24.4 25.7 28.5 25.3 31.1 36.8 28.8 28.2 27.0 30.9 27.1 46.2 39.3

34.2 34.5 44.5 20.7 17.8 16.3 18.0 18.1 15.9 18.8 22.7 18.5 18.6 18.2 20.3 16.1 28.3 23.3

25.9 26.8

36.8 35.2 34.2 43.2 32.1 31.5 36.6 34.9 37.1

21.2 19.9

32.0 22.5 22.5 37.0 36.2 28.4

27.1 19.0 17.3 21.4 19.4 20.7

a a a a a a a a a a a a a a a a a a a a a a a a a a a a a

p p p p p p p p p p p p p p p p p p p p p p p p p p p p p

UMMZ 210347 USNM 559589 KU 85378; KU 85378, MCZ 340296 (s) UMMZ 214229; 214229, ⫺30 (s) KU 85213; UMMZ 214232 (s) KU 43514; KU 41018 (s) UMMZ 214231; 214231, 231601 (s) UMMZ 212919; 217503 (s) UMMZ 158083; 158083, 209198 (s) UMMZ 218545; 209197, 218545 (s) UMMZ 206513; 214226 (s) UMMZ 223664 UMMZ 210982; 207451 (s) UMMZ 228020; 233059, ⫺60 (s) UMMZ 214227; 214227, ⫺28 (s) UMMZ 233061 UMMZ 233058 USNM 557504

FMNH 357420 UMMZ 212917; 212918, 217502 (s) UMMZ 212916 USNM 318240 UMMZ 234936 FMNH 384675 UMMZ 207052; 119430, 151115, 207051 (s) USNM 560668 UMMZ 207450

Behavior, a⫽arboreal; t⫽terrestrial. Breeding behavior, c⫽cooperative; n⫽nesting; p⫽brood-parasitic. 3 First specimen indicates leg measurements, second specimen(s) indicates brain measurements where a specimen other than the first was used, repeated number indicates that brains of all specimens were used. 2

Morphology 31

Eudynamys scolopacea 222 2.59 Urodynamis taitensis 120 2.08 Scythrops novaehollandiae 684 5.55 Chrysococcyx osculans 30 0.77 Chrysococcyx basalis 24 0.56 Chrysococcyx minutillus 18 0.58 Chrysococcyx lucidus 23 0.58 Chrysococcyx caprius 32 0.88 Chrysococcyx klaas 26 0.61 Chrysococcyx cupreus 38 0.94 Cacomantis pallidus 82 1.52 Cacomantis sonneratii 37 0.94 Cacomantis merulinus 24 0.99 Cacomantis variolosus 34 0.93 Cacomantis flabelliformis 44 1.05 Surniculus velutinus 36 0.99 Hierococcyx sparverioides 150 2.19 Hierococcyx hyperythrus 115 1.82 Hierococcyx nisicolor 75 Hierococcyx fugax 80 Hierococcyx pectoralis 76 1.62 Cuculus clamosus 85 1.54 Cuculus solitarius 74 1.59 Cuculus micropterus 119 – Cuculus poliocephalus 52 – Cuculus rochii 59 1.01 Cuculus canorus 106 1.67 Cuculus optatus 117 1.52 Cuculus saturatus 91 – Body weights are from the species accounts.

32 The Cuckoos Brood-parasitic cuckoos have smaller brains for their body size than nesting cuckoos (Table 4.1, Figure 4.1). The variation in brain size is in part explained by a taxon level effect with less scatter around a genus than between genera. Nevertheless, the trend for brood parasites to have smaller brains appears in all three clades of brood-parasitic cuckoos. Cuculus and related brood-parasitic cuckoos have smaller brains than the malkohas, Clamator have smaller brains than the nesting malkohas of the same body size, and the New World brood-parasitic Tapera and Dromococcyx have smaller brains than the nesting Lesser Ground-cuckoo Morococcyx erythropus and Lesser Roadrunner Geococcyx velox. It is unknown whether certain parts in the brain of brood-parasitic cuckoos are small and associated with breeding behavior. In the other brood-parasitic birds that have been examined, cowbirds have larger hippocampus volumes than do the non-parasitic icterids, and females cowbirds that search for host nests have larger hippocampus volumes than males that do not; total brain volume was not determined (Sherry et al. 1993, Reboreda et al. 1996). In birds, the hippocampus is associated with spatial experience and memory (Clayton et al. 1997, Clayton 1998). No other comparative studies on brains of parasitic and non-parasitic birds have been reported. The cooperatively-breeding crotophagine cuckoos do not have large brains for their body size; neither do cooperatively-breeding songbirds have larger brains than other songbirds of the same body size (Iwaniuk and Arnold 2004). In contrast, primate species with the cognitive demands of sociality in complex societies have large brains (Harvey and Krebs 1990, Dunbar 1995, Barton 1996, Kudo and Dunbar 2001). Arboreal cuckoos have smaller brains than terrestrial cuckoos Coua, Carpococcyx and Centropus, in contrast to mammals, where arboreal squirrels have larger brains than terrestrial rodents, as might be expected with the arboreal squirrels’ use of a threedimensional habitat (Eisenberg 1981). Cuckoo brains do not have unusually large olfactory lobes when compared with those of other birds, although the brains of only four cuckoo species have been examined (Bang and Wenzel 1985). The presence of large anal glands in anis (Quay 1967), and

their possible function ( perhaps as in the Hoatzin Opisthocomus hoazin or “Stink-bird” (Newton 1896) which have large olfactory lobes) suggests a direction for further study of olfactory behavior and brain structure in cuckoos. Second, the appearance of the head and bill differs among cuckoos. In the anis the head shape is determined by the bill, which is compressed laterally and expanded into a lateral shield. Bill size and shape may be important in social displays (males have a more prominent profile than females, and adults have a more prominent arc than juveniles), and bills also vary with feeding behavior, with larger bills in species that forage in wet, dense foliage. Bill size and shape vary in the malkohas where bright colors suggest their importance as social signals, and in the lizard-cuckoos where the bill is long and slender. Bills of fruit-eating cuckoos are particularly gross. The nostril shape differs among cuckoos and in early systematic treatments the nostril shape was used to characterize the genera of malkohas (Sharpe 1873).The nostril is circular in many brood-parasitic cuckoos and in the New World cuculines, and slitlike in other brood-parasitic cuckoos. The nostril shape even varies between geographic vicariants within a single species in the Chestnut-breasted Malkoha Phaenicophaeus curvirostris, where the colors of the face and brightly-colored bill with light green, red and black are in different patterns across the range of the species, and the nostril shape varies from a horizontal oval to a vertical slit, with or without a groove extending from the nostril to the cutting edge of the bill. In these birds the nostrils are as variable as the color patterns on the bill, and they may be behavioral signals within a mated pair. In other cuckoos, nostril shape is consistent across species that are closely related in the phylogenetic analyses. Colors of the bill and bare facial skin range from indistinct blackish to bright red, yellow, blue and green. The yellow banana-like bills of some Asian malkohas suggest toucans to a New World ornithologist, and the bicolor huge banana-like bill of Australian Channel-billed Cuckoo led to their original descriptions as “Anomalous Hornbill” and “Psittaceous Hornbill”.A bright blue face is striking both in real life and in photographs of Raffles’s

Morphology 33 Malkoha Rhinortha chlorophaea, Blue-faced Malkoha Phaenicophaeus viridirostris, and several couas Coua, as well as in roadrunners Geococcyx, large forest-living New World ground-cuckoos Neomorphus and Lesser Ground-cuckoo. In some cuckoos the blue contrasts with the neighboring bare skin of orange, red or purple, and in others especially couas the blue is outlined with black. In earlier work on birds, blue skin has often been attributed to the incoherent scattering of light by small particles, the bright violets, blues and greens on the bare faces of these cuckoos may be due to constructive reflection from ordered arrays of collagen arrays, as in other birds with bright blue bare skin. The red facial skin of Red-faced Malkoha is covered with rough papillae, the color disappears when the birds are dead; the source of the color is unknown. Brightly-colored bare skin occurs in many cuckoos, in both the Old World and the New World, especially in the tropical

forests. Bare skin that ranges from violet, blue and green to yellow which reflects certain colors due to constructive interference has evolved at least twice in the cuckoos, once in the Old World and once in the New World. These colors may be associated with colors of the ambient light in forests (Endler 1993, Prum and Torres 2003). Third, the length and proportion of the leg bones vary with arboreal or terrestrial behavior. The legs are longer in proportion to body size in terrestrial cuckoos than in arboreal cuckoos. This morphological adaptation across clades may not have been noted before in birds, perhaps because the mechanics of movement differ between birds and mammals, a group where functional comparisons have been made in more detail (Hildebrand 1974). Engels (1938) and Berger (1952) described the lengths of leg bones in four species of North American cuckoos, but their motive was description

Figure 4.2. Body weight vs total leg length in arboreal and terrestrial cuckoos.

34 The Cuckoos and classification, and not a comparative test of behavior over different clades. In the present study, for cuckoo species that were available as skeletons, the femur, tibiotarsus, and tarsometatarsus were measured and the percentage of the total length of these leg elements was compared with body size (weight) and with terrestrial or arboreal behavior. Across a wide range of body size (23 g to 769 g) the ratios vary with the femur 25–33% of total length, the tibiotarsus 40–44%, and the tarsometatarsus 22–31%.The distal bones do not increase in length disproportionately to the femur with increasing body size, a trend that occurs when birds are compared over a wider range of size (Calder 1984). Terrestrial cuckoos have longer legs and a longer tarsometatarsus than climbing arboreal cuckoos of the same body size (Table 4.1, Figures 4.2, 4.3). Arboreal cuckoos may have short legs to balance and keep a low center of gravity as they move along branches in the canopy. Terrestrial cuckoos with

long legs and a long tarsometatarsus move rapidly without the mechanical problem of high distal mass: angular momentum and the energy to move the legs is an exponential function of the mass of the distal elements, and speed is a function of stride with long legs. Greater Roadrunners on the desert floor for example run very fast with a long stride in pursuit of their lizard prey. The ratios of lengths of leg bones vary less in cuckoos than among other families of birds (Gatesy and Middleton 1997) and large mammals (Scott 1985). Much as in birds, in cursorial ungulates the distal leg bones are longer than the femur.Very large mammals have short distal elements of the leg and a long femur; giants support much weight on their legs, whereas cuckoos are built on a smaller scale. In contrast to the cuckoos, terrestrial squirrels are shorter-legged than arboreal squirrels; the terrestrial squirrels burrow with short legs while arboreal squirrels use their limbs to bound about on branches (Bryant 1945,

Figure 4.3. Body weight vs % tarsometatarsus: leg length in arboreal and terrestrial cuckoos.

Morphology 35 Emry and Thorington 1982, Thorington and Thorington 1989). Cuckoos with rounded wings tend to have a short tarsus or a long digit 3, and there may be other variation in toe length with life style. In all cuckoos, toe 1 is shortest, then toes 2, 4 and 3 (Stephan 2002). In the terrestrial cuckoo that is best known, the Greater Roadrunner, the pelvic muscles originate further anteriorly and laterally than in arboreal

cuckoos. These muscles may support stability and balance in a runner that places its weight alternately on one leg and then the other. Roadrunners have slender leg bones especially the tarsometatarsus, and the toes are mobile and flexible allowing the bird to vary the shape of its foot as it moves on a flat surface (Berger 1952, 1953a, Hughes 1996b). Among the couas Coua of Madagascar, the behavioral differences between species do not appear to co-vary with the dimensions of the tarsus (Milne-Edwards

Table 4.2 Comparative data on the biology of cuckoos, from the species accounts. Species Cuculinae (brood-parasitic) Clamator jacobinus, Jacobin Cuckoo Clamator levaillantii, Levaillant’s Cuckoo Clamator coromandus, Chestnut-winged Cuckoo Clamator glandarius, Great Spotted Cuckoo Pachycoccyx audeberti,Thick-billed Cuckoo Cuculus sparverioides, Large Hawk-cuckoo Cuculus varius, Common Hawk-cuckoo Cuculus vagans, Mustached Hawk-cuckoo Cuculus hyperythrus, Hodgson’s Hawk-cuckoo Cuculus nisicolor,Whistling Hawk-cuckoo Cuculus fugax, Javan Hawk-cuckoo Cuculus pectoralis, Philippine Hawk-cuckoo Cuculus solitarius, Red-chested Cuckoo Cuculus clamosus, Black Cuckoo Cuculus micropterus, Indian Cuckoo Cuculus canorus, Common Cuckoo Cuculus gularis, African Cuckoo Cuculus saturatus, Himalayan Cuckoo Cuculus optatus, Oriental cuckoo Cuculus poliocephalus, Asian Lesser Cuckoo Cuculus rochii, Madagascar Lesser Cuckoo Cuculus mechowi, Dusky Long-tailed Cuckoo Cuculus olivinus, Olive Long-tailed Cuckoo Cuculus montanus, Barred Long-tailed Cuckoo Cacomantis pallidus, Pallid Cuckoo Cacomantis leucolophus,White-crowned Cuckoo Cacomantis sonneratii, Banded Bay Cuckoo Cacomantis merulinus querulus, Plaintive Cuckoo

m wl f wl m wt f wt egg l 154 174 161 199 223 236 201 146 199 178 173 175 172 174 195 221 215 183 206 150 161 138 146 149 193 172 121 110

153 174 160 193 221 227 191 143 196 181 175 174 169 178 194 210 205 175 196 146 160 135 143 143 192 169 124 110

79 90 129 121 77 139 129 92 111 151 101 105 58 62 115 78 76 76 79 78 70 90 87 119 119 117 106 97 110 79 73 99 89 54 54 65 54 58 66 56 59 58 90 86 115 (99) 35 27 25

egg w

egg g

inc da

nestl da

27 26 27 33 24 27 31

22 21 23 24 17 19 20

7.2 6.3 7.9 10.5 3.8 5.4 6.9

12 12

18 13

12.8 13

24 28

28 24

20 16

6.1 3.4

19–20

22 26 25 23 25 21 22 22 18

18 19 19 17 18 16 16 16 14

3.9 12–14 5.2 14 5.0 12 3.6 12 4.5 3.0 3.1 3.1 2.0

17–21 20–21 21 17–20 20–23

23 21 25

16 15 16

3.3 2.6 3.5 12–14

18 20

14 14

2.0 2.2

18–20

36 The Cuckoos Table 4.2 contd. Species Cacomantis variolosus sepulcralis, Brush Cuckoo Cacomantis castaneiventris, Chestnut-breasted Cuckoo Cacomantis flabelliformis, Fan-tailed Cuckoo Chrysococcyx osculans, Black-eared Cuckoo Chrysococcyx basalis, Horsfield’s Bronze-cuckoo Chrysococcyx m. minutillus, Little Bronze-cuckoo Chrysococcyx lucidus plagosus, Shining Bronze-cuckoo Chrysococcyx ruficollis, Rufous-throated Bronze-cuckoo Chrysococcyx meyeri,White-eared Bronze-cuckoo Chrysococcyx maculatus, Asian Emerald Cuckoo Chrysococcyx xanthorhynchus,Violet Cuckoo Chrysococcyx klaas, Klaas’s Cuckoo Chrysococcyx cupreus, African Emerald Cuckoo Chrysococcyx caprius, Diederik Cuckoo Surniculus dicruroides, Fork-tailed Drongo-cuckoo Surniculus velutinus, Philippine Drongo-cuckoo Surniculus lugubris, Square-tailed Drongo-cuckoo Microdynamis parva, Dwarf Koel Eudynamys s. scolopacea, Common Koel Eudynamys s. cyanocephala, Common Koel Urodynamis taitensis, Long-tailed Cuckoo Scythrops novaehollandiae, Channel-billed Cuckoo Cuculinae (Old World), nesting Ceuthmochares australis,Whistling Yellowbill Phaenicophaeus curvirostris, Chestnut-breasted Malkoha Phaenicophaeus pyrrhocephalus, Red-faced Malkoha Phaenicophaeus diardi, Black-bellied Malkoha Phaenicophaeus sumatranus, Chestnut-bellied Malkoha Phaenicophaeus tristis, Green-billed Malkoha Phaenicophaeus viridirostris, Blue-faced Malkoha Taccocua leschenaultii, Sirkeer Malkoha Rhinortha chlorophaea, Raffles’s Malkoha Zanclostomus javanicus, Red-billed Malkoha Rhamphococcyx calyorhynchus,Yellow-billed Malkoha Dasylophus superciliosus, Rough-crested Malkoha Dasylophus cumingi, Scale-feathered Malkoha Cuculinae (New World), nesting Coccycua pumilus, Dwarf Cuckoo Coccycua cinerea, Ash-colored Cuckoo Coccyzus erythropthalmus, Black-billed Cuckoo Coccyzus americanus,Yellow-billed Cuckoo

m wl 116 110 144 118 103 94 106 95 89 109 100 96 104 117 136 117 126 105 195 217 189 354

f wl m wt f wt egg l 115 33 32 20 113 30 32 21 141 50 50 21 117 29 32 20 101 22 24 18 92 19 19 19 106 23 25 18 95 21 24 89 18 21 109 25 30 17 100 20 26 16 95 26 29 19 102 38 37 20 121 34 41 21 135 38 22.5 117 33 38 124 30 31 21 102 43 51 189 229 31 215 262 257 34 186 110 124 24 347 708 660 41

egg w 15 15 15 14 12 13 12

egg g inc da 2.5 12–13 2.6 2.6 2.2 1.4 11–13 1.8 1.4 13–14

12 12 13 15 15 17

1.4 1.3 1.8 11–12 2.5 13–14 2.6 12 3.6

16

3.0

124 169 155 130 139 163 135 158 115 144 177 151 157

124 172 157 131 139 161 131 154 116 145 180 157 159

68 64 30 152 144 39.7 36 66 57 30 114 28.5 116 114 34 84 29 219 140 36 56 62 28 98 97 29 36 115 120 174 167

23 28.4 27 24.5 23.5 26 25 26 24 23 31

102 107 135 139

102 110 139 144

36* 37 47 58

19.6 19 21 23

24.9 45 25.5 54 27 65 30

23 24 17 29

9.1 13–14 10.8 14–17 3.8 16 19.1 8.7 17.7 14.5 10.0 8.7 13.5 10.0 13.5 8.9 8.5 19.1

nestl da 17–19

18 17–19 18–21

19–21 18–20 20–22

19–28 19–28 21–29 17–24

13

5.3 12 5.1 6.6 10–11 8.8 10–11

12 8–9 8–9

Morphology 37 Table 4.2 contd. Coccyzus euleri, Pearly-breasted Cuckoo Coccyzus minor, Mangrove Cuckoo Coccyzus melacoryphus, Dark-billed Cuckoo Coccyzus lansbergi, Gray-capped Cuckoo Coccyzus merlini, Cuban Lizard-cuckoo Coccyzus vieilloti, Puerto Rican Lizard-cuckoo Coccyzus longirostris, Hispaniolan Lizard-cuckoo Coccyzus vetula, Jamaican Lizard-cuckoo Coccyzus pluvialis, Chestnut-bellied Cuckoo Coccyzus rufigularis, Rufous-breasted Cuckoo Piaya cayana, Squirrel Cuckoo Piaya melanogaster, Black-bellied Cuckoo Coccycua minuta, Little Cuckoo Couinae, nesting Carpococcyx radiatus, Bornean Ground-cuckoo Carpococcyx renauldi, Coral-billed Ground-cuckoo Coua gigas, Giant Coua Coua coquereli, Coquerel’s Coua Coua serriana, Red-breasted Coua Coua reynaudii, Red-fronted Coua Coua cursor, Running Coua Coua ruficeps, Red-capped Coua Coua cristata, Crested Coua Coua verreauxi,Verreaux’s Coua Coua caerulea, Blue Coua Centropodinae, nesting Centropus celebensis, Bay Coucal Centropus unirufus, Rufous Coucal Centropus melanops, Black-faced Coucal Centropus nigrorufus, Sunda Coucal Centropus milo, Buff-headed Coucal Centropus goliath, Goliath Coucal Centropus violaceus,Violaceous Coucal Centropus menbeki, Greater Black Coucal Centropus chalybeus, Biak Coucal Centropus ateralbus, Pied Coucal Centropus p. phasianinus, Pheasant Coucal Centropus p. spilopterus, Kai Coucal Centropus bernsteini, Lesser Black Coucal Centropus rectunguis, Short-toed Coucal Centropus steerii, Black-hooded Coucal Centropus s. sinensis, Greater Coucal

128 137 115 114 174 129 134 124 177 171 171 136 105

133 132 116 114 180 130 137 126 194 184 184 135 105

49 64 46 50 134 80 92 95 130 128 104 99 41

57 67 54

31 29 26 40 34 37 32.5 40 38 35 30 24

23 22 20 30 25 25 24 31 25 26 23 19

9.1 7.8 11–12 5.8 27.1 11.8 12.8 10.4 21.3 13.1 13.1 18 8.8 4.8

259 278 221 144 165 137 130 168 139 134 195

253 455(540) 47 278 (400) 44 214 413 420 43.5 148 33.5 168 298 139 240 163 36 133 118 34 168 190 34 146 144 145 35 131 200 236 254 37

35 34 32 25

35.6 28.1 18–19 25.7 11.6

173 156 162 208 272 263 311 225 203 202 226 228 174 164 159 203

182 162 168 201 162 211 238 224 267 769* 271 401 622 261 (500) 221 503 529 203 214 330 342 249 302 445 252 180 146 180 197 238* 164 179 163 215 255 370

155 92 105 189 100 93 40

28 23 28 26.5

15.6 9.9 14.7 13.6

28.5

16.6

31 39

26 31

11.6 21.3

43 37

34 30

27.5 18.4

41 38 36 32 37

33 29 27 25.5 30

24.7 17.7 14.5 11.5 18.4

36

28

7–9 8–13

8⫹

18–24

15

17–21

15.6 15–16

18–22

38 The Cuckoos Table 4.2 contd. Species m wl f wl m wt f wt egg l egg w egg g inc da nestl da Centropus s. andamanensis, Brown Coucal 180 183 234 35 28 15.2 Centropus toulou, Madagascar Coucal 148 168 173 180 33 26 12.3 14–16 19 Centropus grillii, African black Coucal 152 169 100 151 31 24 9.9 18–20 Centropus bengalensis, Lesser Coucal 134 159 86 148 28 24 8.9 Centropus v. viridis, Philippine Coucal 148 160 112 134 30 25 10.4 14 Centropus chlororhynchus, Green-billed Coucal 167 175 35 27 14.1 Centropus leucogaster, Black-throated Coucal 189 198 293 336 Centropus anselli, Gabon Coucal 186 195 210 Centropus monachus, Blue-headed Coucal 187 186 171 237 35 27 14.1 Centropus cupreicaudus, Coppery-tailed Coucal 214 224 272 299 38 28 16.5 17⫹ Centropus senegalensis, Senegal Coucal 160 164 169 169 34 26 12.7 17–19 18–20 Centropus superciliosus,White-browed Coucal 146 154 124 136 33 25 11.4 14–15 18–20 Crotophaginae, nesting (cooperative breeder) Crotophaga major, Greater Ani 203 201 171 152 45 38 35.9 13–14 8–10 Crotophaga ani, Smooth-billed Ani 149 149 111 92 35 27 14.1 13–14 13–17 Crotophaga sulcirostris, Groove-billed Ani 135 131 80 67 31 24 9.9 12–14 10 Guira guira, Guira Cuckoo 179 174 139 145 43 32 2.46 12–13 12–15 Neomorphinae (nesting or brood parasite) Tapera naevia, American Striped Cuckoo (brood parasite) 119 113 49.4 51.1 22 17 3.5 15–16 18 Dromococcyx phasianellus, Pheasant Cuckoo “ 166 163 85 98 25 14.5 2.9 Dromococcyx pavoninus, Pavonine Cuckoo “ 134 137 50 45 21 15 2.6 Morococcyx erythropygus, Lesser Ground-cuckoo 101 102 63 68 27 21 6.6 Geococcyx californianus, Greater Roadrunner 178 171 344 309 39 30 19.4 17–18 17–19 Geococcyx velox, Lesser Roadrunner 146 141 186 174 35 26 13.1 Neomorphus geoffroyi, Rufous-vented Ground-cuckoo 169 164 350 386 43 32 24.3 Neomorphus rufipennis, Rufous-winged Ground-cuckoo 170 164 435 328 40 31 21.3 Neomorphus pucheranii, Red-billed Ground-cuckoo 172 166 330 Data are from the species accounts: m wl ⫽ male wing (mm), f wl ⫽ female wing (mm), m wt ⫽ male weight (g), f wt ⫽ female weight (g), egg l ⫽ egg length (mm), egg w ⫽ egg width (mm), egg g ⫽ egg weight (g), inc da ⫽ incubation period (days), nestl da ⫽ nestling period (days). In Graph 1, for Carpococcyx, brain size is for Carpococcyx renauldi and body size is for C. radiatus; the ground-cuckoos are about the same in other body measurements. Figures in parens were based on single individuals of atypical body condition, as described in the species accounts, and these are not included in all graphs. * indicates unsexed birds.

and Grandidier 1879, Berger 1960); tarsi are long in the arboreal Coua caerulea and the terrestrial C. gigas and C. ruficeps, and short in the climber C. reynaudii and the terrestrial C. cursor. The main difference between couas is in the longer tendons and the site of their attachment in the long-legged forms

(Berger 1960). Fourth, wing shape in cuckoos varies with the longest primary ranging from the 9th (the most pointed wing) to the 2nd (the most rounded and broad wing).The wing length and shape vary with migratory behavior. Pointed wings are seen in the

Morphology 39

Figure 4.4. Wing length and body mass in cuckoos.

migratory Cuculus species and rounded wings in the ground-cuckoos and in coucals Centropus, where the longest primary is P7, P6 or P5 in Lesser Coucal C. bengalensis and P4, P3 or P2 in the large Greater Black Coucal C. menbeki. The migratory northern hawk-cuckoos have more pointed wings than the more tropical species (Rufous Hawkcuckoo Hierococcyx hyperythrus and Whistling Hawk-cuckoo H. nisicolor, compared with Javan Hawk-cuckoo H. fugax and Philippine Hawkcuckoo H. pectoralis). The long wings of broodparasitic cuckoos in the Old World are related to their migratory behavior and not to brood parasitism, insofar as the New World brood-parasitic cuckoos are residents and have short wings (Table 4.2, Figure 4.4). Many nesting cuckoos are shortwinged, especially resident tropical birds, and so are a few forest brood parasites especially the African long-tailed cuckoos Cercococcyx, which have a

rounded wing with the inner primaries P4-7 longer than the outer primaries P8-9. Wing shape in cuckoos has been described in detail from the lengths of the wing feathers in relation to ulna length (Stephan 2001a, Figure 4.5). In Figure 4.5, the primaries are numbered from the proximal end of the carpometacarpus, the secondaries S1 from the distal end of the ulna, S6 and the more inner secondaries from the proximal end of the ulna, and S2-5 from their corresponding points of attachment along the ulna. In another comparison, the length of “free” primaries relative to “free” secondaries captures the shape of the wing, the term indicating the length of feathers that extend beyond the wing coverts (Stephan 2001a, Figure 4.6). All cuckoos can fly and have a keeled sternum, even in the terrestrial roadrunners which more often walk and run than fly. The shape of the

Figure 4.5. Wing shape in cuckoos, with relative lengths of the primaries and secondaries (after Stephan 2001a). Lengths of the wing feathers are plotted from their positions of attachment along the ulna, which is represented by the two small circles on each wing: a, Scythrops novaehollandiae; b, Cuculus canorus; c, Eudynamys scolopacea malayana; d, Clamator levaillantii; e, Rhamphococcyx calyorhynchus; f, Piaya cayana; g, Centropus celebensis; h, Coua ruficeps; i, Tapera naevia; j, Geococcyx californianus.

Morphology 41

Figure 4.6. Wing shape in cuckoos, with extent of “free” primaries and secondaries: a, rounded wing, Geococcyx californianus; b, pointed wing, migratory Cuculus canorus (after Stephan 2001a).

sternum varies with life habits among Coua species, the more arboreal couas with a more convex and strongly keeled sternum than the terrestrial couas (Milne Edwards and Grandidier 1879). Finally, much variation in wing morphology of cuckoos can be explained by their phylogenetic associations, such as coucals being short-winged owing to being related to other short-winged coucals, and secondarily owing to their terrestrial and non-migratory behavior. Coucals have short wing bones as well (Milne-Edwards and Grandidier 1879, Berger 1952, 1953a); and the New World terrestrial roadrunners Geococcyx and to a lesser extent the anis Crotophaga have shorter wing bones relative to leg bones than the Coccyzus cuckoos (Larson 1930, Engels 1938, Berger 1952).

Sexual size dimorphism Among most cuckoos the sexes are nearly the same in size (Table 4.2, Figures 4.7, 4.8). Sexual dimorphism in body size does not vary among cuckoos as

much in relation to their breeding biology as to their systematic status. Among the coucals, females are larger than males in the monogamous species such as Pheasant Coucal, where the female is half again the mass of the male and both sexes care for the young, and in the sometimes-polyandrous African Black Coucal where the male looks after the young. The degree of sexual size dimorphism is consistent across subspecies in the coucal species that vary geographically in size (as in Greater Coucal and Philippine Coucal). The largest degree of sexual size dimorphism occurs in the smaller coucals, unlike birds such as grouse where sexual size dimorphism is greater in the larger species (Rensch 1950, Payne 1984, Abouheif and Fairbairn 1997). Among the smaller coucals the females are larger than the males, with larger coucals the sexes are the same size, and with the largest coucals the males are larger than the females. With other kinds of birds, reversed sexual size dimorphism where females are the larger sex is sometimes associated with polyandry, as in buttonquail, tinamous Tinamidae, painted snipe Rostratulidae, jacanas Jacanidae and some other shorebirds (Payne 1984, Andersson 1994, 1995). Insofar as reversed sexual size dimorphism occurs both in some monogamous birds and in polyandrous birds, the large size of female coucals does not necessarily indicate a polyandrous mating system among coucals. Other explanations of sexual size dimorphism include the ability of large females to lay more or larger eggs, or to defend the young where the small and agile males forage for the females and leave them at the nest, as with many monogamous hawks and owls. Reduced competition for food between mates among some birds is also a correlate of sexual size dimorphism, though the use of prey of different sizes may follow sexual size dimorphism that evolved for other reasons (Andersson 1994). Coucals have a long hallux claw, much as do some other terrestrial birds such as larks and pipits, while other birds have long claws on all the toes (longclaws Macronyx spp., jacanas Jacanidae, Longtoed Lapwing Vanellus crassirostris, Purple Heron Ardea purpurea), birds that live in grassland and on weak-stemmed and floating aquatic vegetation. In coucals the hallux claw ranges from 32–35% of tarsus length (in the large Buff-headed Coucal

42 The Cuckoos

Figure 4.7. Sexual size dimorphism in body size in cuckoos: body mass.

Centropus milo and Violaceous Coucal C. violaceus), to 68–76% of tarsus length (in the small marshliving African Black Coucal C. grillii and Lesser Coucal C. bengalensis). An exception is Short-toed Coucal C. rectunguis in which the hallux claw is only 23% of tarsus length. A test of whether the hallux claw is a sexually selected trait in coucals is whether the claw is proportionally larger in one sex. The diagonal line in Figure 4.9 indicates the position where male and female claws are equally long in relation to tarsus length. About the same number of coucal species have males with relatively longer claws on average than females. When claw length is compared with body size, there is no trend for species with a large size to have proportionally larger claws in females than in males. Also there is no increase in relative size of the claw of the sex with larger body size, as there is with measurements in other polygamous birds (Payne 1984).The lack of sexual

dimorphism in this prominent feature of coucals does not support the idea that the reversed sexual size dimorphism of coucals has been driven by polyandry. Sexual differences in size of cuckoos might be related to female aggressive behavior. In some hawks, falcons and owls, the female takes the lead in social dominance or territorial behavior of the pair (Pleasants and Pleasants 1989, Mueller 1990). Field observations on mating among cuckoos do not suggest female social dominance over males. In the cooperatively breeding anis, females are about the same size as males. Females compete for the position of their eggs in a communal nest and they remove each others’ eggs, yet within a group the body size is not larger in the socially dominant females at a nest (Vehrencamp et al. 1986). Large female size may also be related to egg size in anis, where large eggs provide nestlings a head start for growth in the competition for parental care in a brood.

Morphology 43

Figure 4.8. Sexual size dimorphism in body size in cuckoos: wing length.

Plumage The body plumage of ground-cuckoos, couas and malkohas is soft and lax.The plumage becomes wet in the rain, and the birds often sun themselves to dry after a shower and even on a warm morning (Forbush 1927, Cracraft 1964). The anis have smooth, dense glossy plumage, a texture that appears to be related to foraging in wet foliage. Their feathers still become soaked, and in cool wet weather the anis spread their wings in the sun’s rays to dry (Skutch 1983), like a vulture or a cormorant. Coucals that forage in wet habitats do the same. None of these plumage traits are exclusive to cuckoos, as some gruiform birds and doves have the same kind of lax plumage and soft pastel colors of purples and pinks like the plumage of couas and malkohas. Cuckoos also have bristles, specialized contour feathers with a stiff rachis, and barbs are unfused and restricted to the feather base (Stettenheim 1973).

Specialized feathers are found on the face of many nesting cuckoos. The eyelashes are small and inconspicuous in several cuckoos but are conspicuous, stiff and curved without vanes or barbs in some malkohas (Yellow-billed Malkoha Rhamphococcyx calyorhynchus, Green-billed Malkoha Phaenicophaeus tristis, Black-bellied Malkoha P. diardi, short and thin in Blue-faced Malkoha P. viridirostris, Rough-crested Malkoha Dasylophus superciliosus and Scale-feathered Malkoha D. cumingi), and they are particularly long and stout in Sirkeer Malkoha Taccocua leschenaultii which live in semi-arid sandy habitats.The eyelashes are short in Red-faced Malkoha P. pyrrhocephalus and in the African yellowbills Ceuthmochares; they are absent in the small Raffles’s Malkoha Rhinortha chlorophaea. The eyelashes are conspicuous in New World squirrel-cuckoos Piaya; they are short in Coccycua and Coccyzus. Eyelashes are also found in anis and in large terrestrial cuckoos including Coua, Carpococcyx, Centropus, Geococcyx and Neomorphus. Eyelashes are short in Clamator and are absent or

44 The Cuckoos

Figure 4.9. Size and sexual size dimorphism in hallux claw in coucals Centropus. Measurements are in the species accounts.

not noticeable in other Old World brood-parasitic cuckoos. The face and head have some feathers with long delicate projections of the rachis or shaft from the loose-webbed base of vaned feathers in the malkohas. In Sirkeer Malkoha these shafts are stiff and only a small length of the feather base has a vane. In Chestnut-bellied Malkoha P. sumatranus the feathers are recurved forward especially around the base of the bill. In Blue-faced Malkoha P. viridirostris the lower head, neck and breast feathers are forked. In these unique feathers, the bifurcation is formed by stiff terminal barbs that cling together on each side of the central feather shaft, which protrudes as a fine free black hair-like bristle. Delicate hair-like shafts occur in some other cuckoos including Great Spotted Cuckoo, although these structures are not evident in most Old World brood-parasitic cuckoos

(Cuculus, Chrysococcyx, Eudynamys, Scythrops). In the New World, the Guira Cuckoo and roadrunners have long delicate feathers around the face (Figure 4.10). The feathers on the face and head occur in an extreme form with a thickening and fusion of barbs in two Philippine malkohas. Rough-crested Malkoha Dasylophus superciliosus has superciliary crests formed by stiff reddish barbs; at the base a few more basal barbs form a loose-webbed vane on each side of the shaft. Perhaps the most bizarre head feathers are in Scale-feathered Malkoha D. cumingi whose terminal barbs are fused at the tip into a single shiny metallic flat blue-black shield that contrasts with the white vane near the base of the feather.The tips are formed of fused barbs that are densely packed with melanin, while the surfaces are thick layers of keratin with little melanin (Brush

Morphology 45

Figure 4.10. Facial feathers of cuckoos: Greater Roadrunner Geococcyx californianus and Guira Cuckoo Guira guira.

1965). The tips of the crest feathers are twisted to reveal the same iridescent color on each side of the head. Some coucals have hackles, robust elaborate feathers on the neck, with conspicuous stiff feather shafts, the shaft color often in contrast to the feather vane. In Greater Roadrunner the head has stout, bare, black terminal shafts that project outward and backward over the eyes, barbed bristles around the base of the bill, and short recurved feather shafts all around the head, while Rufousvented Ground-cuckoo has dark bare eyelash bristles above and below the eye, and semibristles in front of the eye. Many cuckoos have long tails that function as rudders in slow flight in the arboreal species, and as steering devices and balancers in the cursorial ground-living species. Cuckoo tails are also conspicuous in courtship behavior.The tails often have

contrasting colors, with broad or narrow white tips in many malkohas and New World arboreal cuckoos, and red tips on green tails in Chestnut-breasted Malkoha P. curvirostris. Raffles’s Malkohas Rhinortha chlorophaea are sexually dimorphic, the males with the tail finely barred black and gray in a washboard pattern and broadly tipped white; the females with the tail rufous and a black subterminal band and white tip. In several cuckoos (Common Cuckoo, Sirkeer Malkoha,Yellow-billed Cuckoo) the tail is spread and held above the back, exposing the color pattern on the dorsal surface in frontal display to the mate. In some cuckoos the edges of the tail feathers are “notched” with a V-shaped area of contrasting color, the apex toward the feather shaft; in others the notches are extended well towards the shaft and almost form a bar. The notches are distinctive of certain species especially of the brush cuckoos Cacomantis. Tails are most elaborate in the small ground cuckoos Dromococcyx of the New World tropics, with the long rectrices arrayed in a fan and the upper tail coverts cascading nearly the full length of the tail, each pointed covert feather with a black band and a delicate white spot at the tip. In display the bird arches the tail coverts over the body.

Natal down and naked nestlings Natal down of cuckoos is hair-like, a single keratinized shaft called a “trichoptile” (hair-feather), rather than an array of non-interlocking barbs as is the fluffy natal down of many kinds of birds.These trichoptiles are not down-like in structure but are down-like in being the first set of feathers on a young bird.They are present at hatching or within a day or two in most cuckoos, and are developmentally homologous with the natal down of other birds. The trichoptile grows from the same follicle as the juvenile contoured feather, and when the pinfeather bursts open, the hair-like natal down remains attached to the tip for a day or two in the same position on the tip of the feather as the fluffy down of songbirds. The down is often worn away by the time the nestling fledges. This down is white in many cuckoos; it differs in color between certain closely-related species (gray in

46 The Cuckoos Yellow-billed Cuckoo, white in Black-billed Cuckoo).A few nesting cuckoos have not been studied as hatchlings but have soft body feathers as feathered nestlings (Scale-feathered Malkoha); these may be loose-webbed juvenile contoured feathers rather than natal down. Malkohas, of which the nestling has been studied and described, have hair-like natal down. In young cuckoo nestlings with dark skin, such as Black-billed Cuckoo and Greater Roadrunner, the white hair-like down forms a strong pattern of contrast with the skin. In the nesting cuckoos, these feathers appear to be continuous with the sheath that covers the growing first generation of contoured feathers. Hair-like down is prominent in the coucals Centropus, where the feathers are white, stiff and sometimes as long as 20 mm and the nestlings look like spiny hedgehogs; less typical of young coucals is Lesser Black Coucal C. bernsteini which has only short white tips attached to the growing contoured feathers on the head. Natal down is absent in the nesting Old World ground-cuckoos Carpococcyx and Madagascar couas Coua, both as hatchlings and when they fledge. In adult cuckoos, fluffy down-like feathers are limited to the apteria. Natal down is absent in the brood-parasitic cuckoos, and this may be explained by the fact that the naked skin of nestlings is sensitive to the touch of host eggs which they evict, and natal down would interfere with this close contact. An exception is Shining Bronze-cuckoo, which has a few small hairlike feathers in New Zealand and New Caledonia, and few or none in southern and Western Australia where the young are naked; Little Bronze-cuckoo also has some natal down. Nestling brood-parasitic cuckoos evict the host egg from the nest; natal down in those bronze cuckoos is mainly on the head, thigh and side, rather than on the back where the egg is balanced in eviction behavior. Nestling crested cuckoos Clamator, like other Old World brood-parasitic cuckoos, lack natal down, and nestling brood-parasitic American Striped Cuckoos in the New World also hatch naked without down.

Plumage change with age Juvenile plumages are variable in a few broodparasitic cuckoos. Jacobin Cuckoo and Levaillant’s

Cuckoo juveniles occur that are dull versions of the adult plumage phases, and in East-Central Africa a rufous color phase occurs in Levaillant’s Cuckoo. Juvenile Diederik Cuckoos vary in plumage in some regions, with rufous juveniles in areas where the cuckoos parasitize bishop finches, Euplectes species. It is unknown whether the color variants of cuckoos are genetically different (diederiks have egg races in these areas, and the plumage may correspond with egg races), or the rufous plumage is a result of bishop foster parents rearing the young on seeds rather than entirely on insects as in the greenplumaged juveniles that are reared by the more insectivorous weavers Ploceus. Juvenile plumage in most cuckoos is similar to adult plumage, especially so in the nesting cuckoos. The transition from juvenile to adult appearance occurs early in life. The main difference in plumage is the shape of the rectrices (narrow and pointed in juvenile plumage, more broad and rounded or truncate in the adult) and the presence and extent of white tips on the rectrices (less in juveniles), or the presence of barred rectrices (barred in juvenile coucals, not barred in adult coucals). Juveniles are quick to assume adult head plumage. Philippine Red-crested Malkoha and Scale-feathered Malkoha with their elaborate head plumage grow these feathers within a few weeks of leaving the nest and have their showy adult head plumage by the time the tail is fully grown. A few flight feathers in wing and tail are often retained into the first breeding plumage of the coucals Centropus spp. The cuckoos with the most strongly age-graded plumage, where juvenile plumage differs from adult plumage, are certain brood-parasites. None of these cuckoos mimic the plumages of their host species. Juvenile Common Koels Eudynamys scolopacea have black plumage like their crow hosts in much of their range, yet this is characteristic only of male juveniles; female koels have a spotted brown juvenile plumage that is intermediate between that of juvenile males and adult females. Behavioral associations with the host species may nevertheless have some effect on juvenile plumage. In koels the plumage somewhat resembles that of their host species, dark like crows in India where

Morphology 47 koels parasitize crows, and streaked brown in Australia where they parasitize brown honeyeaters. The juvenile plumage of Philippine Drongocuckoo Surniculus velutinus is similar to that of a paradise flycatcher Terpsiphone cinnamomea, but the host of this cuckoo is unknown and the similarity of juvenile plumages in these birds is no more than a teasing clue to the mystery of their breeding behavior. A distinctive subadult plumage that is retained for a year after the molt from juvenile plumage occurs in the Asian hawk-cuckoos, as described for some of these birds by Mayr (1938b); previously these had been confused as a female plumage (Siebers 1930). Species with a distinct subadult plumage are Large Hawk-cuckoo Hierococcyx sparverioides, Dark Hawk-cuckoo H. bocki, Rufous Hawk-cuckoo H. hyperythrus, Whistling Hawkcuckoo H. nisicolor, Javan Hawk-cuckoo H. fugax, and Philippine Hawk-cuckoo H. pectoralis. At least one hawk-cuckoo breeds in subadult plumage, Philippine Hawk-cuckoo females laying in this plumage as well as in the definitive unstreaked adult plumage. In the Sulawesi Cuckoo Cuculus crassirostris also a distinct subadult plumage occurs and specimens with this plumage are more common in museum collections than are the juveniles, which have a white head much as that of the closely related Indian Cuckoo C. micropterus, but nothing is known of their breeding. Burchell’s Coucal Centropus superciliosus burchellii breed in subadult plumage as well as in the definitive plumage that lacks the white eye-stripe. Another coucal with subadult plumage is a subspecies of Blue-headed Coucal C. monachus heuglini in the southern Sudan, where its breeding has not been studied. In known cases, the development of sexual maturity is completed well before the development of adult plumage. Among these cuckoos the sexes are alike in plumage and both sexes have the full set of plumages. In both Centropus superciliosus and C. monachus, the other subspecies lack the extra plumage and breed in the plumage that looks like the “subadult” plumage of these two subspecies. The lack of a subadult plumage is probably related to the rapid development of sexual maturity of cuckoos in general.

Cuckoos sometimes retain a few juvenile feathers through their first year, particularly in the wing (e.g., Great Spotted Cuckoo, Diederik Cuckoo Chrysococcyx caprius and Klaas’s Cuckoo Chrysococcyx klaas).The retained juvenile feathers show that these cuckoos breed in their first year, as these birds with mixed plumage have been taken with eggs in the oviduct. Earlier studies misinterpreted the great variation in adult plumage of some cuckoos (e.g., Black Cuckoo, Bannerman 1921) to be due to the number of years it takes a bird to develop an adult plumage. The larger number of specimens now available for examination in museums shows that each adult plumage form follows a single molt from the all-black plumage of the juvenile, in which the tail tip is black and not white as in the adult. Insofar as cuckoos usually breed in their first year, there is no reason to expect a series of subadult plumages. If birds that attempt to breed in their first year are less successful in reproduction over their lifetimes than those that do not attempt to breed at that time, then a delay in sexual activity in the first year may be the more productive life history strategy, and we should see sexual inactivity in birds with a lessthan-definitive plumage. The idea of a “delayed plumage maturation” is derived from the more general concept of “delayed maturation”, which proposes that few first-year birds are competent to breed, and that early reproduction negatively affects their later success (Lack 1954, Selander 1965,Williams 1966).A logical consequence of this life history theory is that individuals with plumages unlike the older adults are sexually inactive. If this is true, we predict that these birds do not breed, or when they do breed they have a lower survival and lifetime reproduction. However in many cuckoos the birds with odd plumages are sexually active. Female cuckoos with a few juvenile feathers in their plumage have been taken with eggs in the oviduct so they were actively breeding, as described in the species accounts of Plaintive Cuckoo C. merulinus, Brush Cuckoo and black-billed Common Koel Eudynamys scolopacea melanorhyncha. There is no indication that brood-parasitic birds have a long life (Payne 1977b).They are expected to breed in their first year: if they deferred breeding in the first year, they would have a high chance of dying before the next year. In some cooperatively-living species of

48 The Cuckoos songbirds the definitive adult plumage may not occur while the young are living in a group with older, socially and sexually dominant individuals (Rowley and Russell 1997). These long-lived cooperatively breeding birds are one group where “delayed maturation” might occur, although with removal of older birds in the group, the younger birds become breeders themselves. However, in the cooperativelybreeding anis Crotophaga and Guira Cuckoo Guira guira which compete for social status at communal nests, where increased maturity might give the birds some benefit in competing with their group members, the species do not have distinct subadult plumages, and these birds do not have a long life span. A few coucals molt into a winter non-breeding plumage before they molt into the adult breeding plumage (African Black Coucal Centropus grillii, Lesser Coucal C. bengalensis, Australian Pheasant Coucal C. p. phasianinus, and some other coucals). These birds are known to molt from breeding plumage back to a nonbreeding plumage, at least in forms where specimens are available from all seasons; a second- and later-year non-breeding plumage is not known for the New Guinea races of Pheasant Coucal. In Blue-headed Coucal Centropus monachus fischeri and in White-browed Coucal C. superciliosus burchellii and C. s. fasciipygialis, a distinctive subadult plumage is retained for a year after the molt from juvenile plumage and before the molt to definitive adult plumage. In addition, Stresemann (1913a) thought Greater Coucal C. sinensis had a series of subadult plumages based on the variation in the retained juvenile feathers in birds in partial adult plumage. He also noted the absence of museum specimens of Lesser Coucal in molt from adult breeding plumage to nonbreeding plumage.The lack of molting specimens may be due to the secretive behavior of birds after the breeding season, rather than to a lack of a post-breeding molt. It would be useful to keep these birds in captivity and observe their sequence of molts and plumages. Some coucals also have a distinct adult nonbreeding plumage, at least in their first year. They molt from juvenile plumage into this alternate nonbreeding plumage and maintain the plumage for a season before they molt into an adult breeding plumage. The nonbreeding plumage has sometimes been

overlooked owing to its similarity with the juvenile plumage. Although there is no evidence of an adult nonbreeding plumage for most coucals, distinct nonbreeding plumages are known for a few species including Lesser Coucal C. bengalensis throughout its range and the Pheasant Coucals C. phasianinus nigricans, C. p. thierfeldi and C. p. spilopterus in New Guinea and the Kai Islands. The question remains whether birds with a distinct subadult or nonbreeding plumage maintain the adult breeding plumage all year into later years, or assume a distinct non-breeding plumage as in birds in molt from black to pale plumage in Pheasant Coucals of Australia, Lesser Coucal, African Black Coucal and Madagascar Coucal C. toulou. Again, it would be informative to watch the plumage change through the seasons within an individual bird species.

Sexual dimorphism in plumage A few cuckoos are sexually dimorphic in plumage color, perhaps more among the Old World parasitic cuckoos where males are conspicuous and glossy and females are cryptic and not so glossy like the African and Asian Chrysococcyx, and the Common Koel where males are black and females are brown. The eye-ring and iris color and pattern (a ring in the iris) also differ between the sexes in some cuckoos, both in glossy cuckoos where sex is apparent from eye color like the Little Bronze-cuckoo and Shining Bronze-cuckoo, and in several malkohas. Conspicuous plumage dimorphism is not known in nesting cuckoos except in Raffles’s Malkoha Rhinortha chlorophaea.

Plumage phases and color morphs Adult plumage is typically cryptically colored in the brood-parasitic cuckoos. In several brood-parasitic cuckoos more than one adult plumage phase occurs within a population and these color phases are limited to the females. In several Cuculus species the females occur in either a gray phase like the adult male or a barred rufous (or “hepatic”) color phase. The proportion of female Common Cuckoos with gray or rufous plumage varies in Europe, where rufous females are most common in Central Europe,

Morphology 49 and the proportion of color phases parallels that of Accipiter hawks and Falco falcons (Voipio 1953).The variation in plumage may be density-dependent, where a bird in the rarer plumage phase has an advantage in not being recognized by naive songbirds, while the more common form is recognized and the cuckoo is chased away from the nest before it can lay (Payne 1967). Distinct plumage color phases (gray without bars, rufous with bars) also occur in breeding female Plaintive Cuckoo and Brush Cuckoo (not in all subspecies). Plumage variation in the Pallid Cuckoo Cacomantis pallidus of Australia appears to the number of rufous feathers grown after the postjuvenile molt (not retained juvenile feathers, which are white and black) and after adult molt in later years in breeding females. Finally, female Common Koels Eudynamys scolopacea melanocephala in Sulawesi have three plumage phases, and females in each of these plumages are known to lay as they have each been collected with an egg in the oviduct (Stresemann 1940). Two species of brood-parasitic crested cuckoos Clamator, Jacobin Cuckoo and Levaillant’s Cuckoo, have color phases in both sexes: a pied phase that is white or whitish below, and a black phase where white is limited to a wing patch (and in Levaillant’s Cuckoo to the tip of the tail). In both cuckoos the black phase is limited to a small part of the distributional range of the species in Africa. Color morphs and locally distinct subspecies that differ in plumage color occur among several nesting cuckoos, particularly the coucals. In some the plumage is partly albinistic. In the Andaman Islands, Greater Coucal Centropus sinensis andamanensis has a pale body (buffy white or dusky gray) rather than a black body as in most other subspecies, and on Kangean I south of Borneo the form C. s. kangeangensis likewise has a pale body (buffy white or dusky gray). On Timor the only specimen collected of Pheasant Coucal C. phasianinus has been described as a subspecies (C. p. mui) and has a white breast, in contrast to other populations of Pheasant Coucal which have black underparts in breeding plumage and buff with paler streaks in nonbreeding plumage. Goliath Coucal C. goliath in the northern Moluccas has a whitish morph in Halmahera. Melanistic or black plumages occur in certain coucal populations on

islands, especially in Philippine Coucal C. viridis with two subspecies having black wings (C. v. mindoroensis, C. v. carpenteri) unlike the rufous wings of coucals elsewhere in the Philippines, and on Luzon an uncommon white phase occurs in a population where most birds are black and rufous. Pied Coucal C. ateralbus in the Bismarck Archipelago have four distinct color phases in the amount of white in the adult plumage; birds with intermediate plumage are also known. Senegal Coucal C. senegalensis have a dark rufous color phase “epomidis” in humid areas of coastal West Africa; other birds in these populations are white below.

Other comments on plumage The cuckoo wing is eutaxic, with eutaxy the number of secondaries equal to the number of upper greater secondary coverts (Stephan 1970). The alternative condition in birds is diastataxy, where the wing has a missing secondary S5 (or an extra secondary covert between S4 and S5). The condition varies among orders of birds and within some families such as the megapodes Megapodiidae, kingfishers Alcedinidae, rails Rallidae and waders Scolopacidae, and it varies among the birds suggested as relatives of cuckoos (diastataxic in turacos Musophagidae, parrots Psittacidae, doves Columbidae and nightjars Caprimulgidae; and eutaxic in trogons Trogonidae, woodpeckers Picidae and the perching birds Passeriformes). Because the covert varies in no apparent regular manner among orders, and it varies within some families (the kingfishers, the waders Scolopacidae), the eutaxic wing gives us no clue to the relationship of cuckoos with other orders of birds (Sibley and Ahlquist 1990). Total plumage mass of Smooth-billed Anis has been measured. Males have an average plumage mass of 4.6 g, and a feathered body of 129.1 g; plumage is 3.6% of the total body weight. Females have an average plumage of 3.9 g and a feathered body of 120.0 g, and plumage is 3.3% of total body mass (Mitsch 1974).This proportion is about average among birds of their body size.

Glossy plumage The glossy- and bronze-cuckoos Chrysococcyx have a glossy plumage, with considerable variation

50 The Cuckoos among the species in color and brilliance. Two species are named for their brilliant greens, Asian Emerald Cuckoo C. maculatus and African Emerald Cuckoo C. cupreus; other cuckoos are named for a coppery or subtle bronze copper such as the Australian Horsfield’s Bronze-cuckoo. The glossy plumage of Asian Violet Cuckoo C. xanthorhynchus is violet purple in mainland Asia and the Sundas, and blue in the Philippines subspecies. Plumage gloss is more intense in the sexually dimorphic African glossy cuckoos and less intense in the sexually monochromatic Australian glossy cuckoos, where only the colors of the bare eye-ring and iris give visual cues to the sex of a bird. There may be an evolutionary conflict in the plumage appearance of these brood-parasites, between sexual selection within the species and avoiding detection by the host (Payne 1967, Andersson 1994). Structural features of feathers affect plumage color, including the wavelength λ of reflected light, the proportion of light that is reflected and the brightness of plumage, and the consistent color of reflected light viewed from different angles. The shape and curvature of the barbules, the changing dimension of barbules along the feather shaft and the orientation of barbules all combine to maintain a constant λ over a wide range of angles of incident light (Durrer and Villiger 1970, Dyck 1987). The constancy of reflected color in plumage when seen at different angles to the incident light in the emerald cuckoos C. maculatus and C. cupreus may result from the rounded shape of the reflectors and the angle of barbules, with the proximal and distal barbules nearly at right angles to each other (Dyck 1987). The structure of body feathers of African Emerald Cuckoo has been examined by light microscopy (Dorst 1951) and scanning and transmission electron microscopy (TEM) (Durrer and Villiger 1970, Durrer 1986). Feather barbules originate on each side of the barb shaft.The barbules are broad and nearly flat along their length, with haemuli (small hooks) on the medial side of the vane, and the barbules are elongated, narrow and lack haemuli on the outer side, the feather structure giving the plumage a frosted appearance. The frosted plumage of this cuckoo contrasts with the more

glossy Diederik Cuckoo and Klaas’s Cuckoo which have only the elongated, unflattened barbules (van Someren 1925, Dorst 1951), so the appearance of texture is due to the microstructure which exposes the color-producing portions of the barbules. Internally the barbule has layers of melanin granules, rod-like, arrayed side by side with the long axis parallel to the long axis of the barbule. The convex upper surface of the barbules has 10–14 of these melanin layers and the inner surface has 5–6 layers separated by keratin, the upper and lower stacks arranged around a keratin core.The number of these layers is related to plumage brightness (Rütschke 1966); glossy cuckoo feathers are more brilliant on the outer surface which reflects light to other birds than on the inner surface which faces the bird’s body, and the Scale-feathered Malkoha has equal brilliance on both sides. Multiple layers of melanin and keratin in the feather barbule are reflectors that determine the constructive interference of light waves. Some light is reflected from the first layer, some from the second, and so on.The color observed is a summation of light waves reflected off each layer. The light reflected from each layer travels a different distance to any point in the feather, and the most reflected waves pass out of phase with one another and cancel each other out. But those waves that travel an entire wavelength in the added distance traveled, are in phase with each other. These waves sum to produce the more brilliant color of a restricted wavelength. In the plumage of other green birds that have been studied in more detail, the structural determinants of reflected color include the number of melanin granule layers (the more layers, the more total light is reflected and the more brilliant the color), and the thickness and number of layers of melanin and keratin (Dyck 1987). Insofar as reflected light waves interfere with light waves reflected from other surfaces and the interference depends on their phase relationships, the wave lengths amplified in constructive interference can be calculated from a physical model of reflected light (Durrer 1986). The model predicts that the optical thickness (actual thickness⫻ refractive index) of each layer is equal to a quarter of the wavelength λ of maximum reflectance. This is the difference in

Morphology 51 light paths (2 ⫻ sine of the angle of incidence ⫻ the distance between reflecting layers) and the refractive indices of melanin and keratin. Reflected color varies with the angle of incidence and observation of light. At the larger critical angle the melanin layers absorb nearly all light, and at an angle of 0o no light would undergo constructive interference and the reflected colors would be less intense than when light is reflected at an angle, but the bigger effect is on hue. Differences between path length of waves become larger as incident light is more direct, and the hue will vary from a shorter wavelength (more blue) to a longer wavelength (more red). In this model of constructive interference, the wavelength λ amplified in an ideal regular multilayer reflector is four times the thickness of the semitransparent layers, times the refractive index. For an African Emerald Cuckoo feather, the λ observed is close to the λ value calculated from the thickness of keratin layers in the feather (Durrer and Villiger 1970). The exact color that is most strongly amplified depends on how the keratin layers are measured in relation to the rounded melanin rods (Prum et al. 1999b). In addition to this ideal model of multiple reflectors, the arrangement of keratin and melanin may affect the reflected colors through “chirping” series of layers where the peripheral reflectors are thick and the interior layers are progressively thinner. Parker et al. (1998) found these chirping multiple reflectors in the glossy gold wings of beetles. Feather barbs of African Emerald Cuckoo have thicker peripheral layers than internal layers of keratin, and TEM figures show a 20–30% shift in keratin thickness between layers (Durrer and Villiger 1970), much as in beetle wings, and this ordered structure may affect plumage color. The layers of different thicknesses may reinforce different components of the spectrum, producing complex and broad spectrum colors, metallic hues with coppery or golden highlights. In addition to the constructive interference of colors that is produced by multiple reflecting layers, coherently scattering threedimensional nanostructures occur in feathers (Prum et al. 1998, 1999a). In other birds with glossy green feathers, the colorful green Malachite Sunbird Nectarinia famosa has

more layers of melanin and keratin in the feather than does the Black Sunbird N. amethystina, while the non-glossy Gray Sunbird N. veroxi has unordered melanin granules that are not in regular layers as in the glossy species (Farquhar et al. 1996). In African starlings there is a similar difference between the dull and modestly glossy birds, but the most glossy birds have oblong melanin granules that are filled with a refractive material, especially the Emerald Starling Coccycolius iris with multiple layers on the dorsal feather surface and a single layer on the ventral surface.The color and gloss of plumage varies with the size and shape of these granules (Craig and Hartley 1985, Durrer 1986). These structural features may determine differences in plumage color among the glossy cuckoo species. The iridescent color of plumage has been said to change with wear, as in Greater Black Coucal Centropus menbeki becoming less blue and more green (Rand 1942a). The mechanism of color reflectance described for the internal layers of melanin and keratin within the feather barb make this kind of change unlikely, and the observation of variation in color probably refers to natural variation in plumage color among birds within a population.

Molt Most cuckoos molt their plumage once a year. Exceptions are coucals which molt their body plumage twice a year, and anis which have a complete molt and an incomplete molt within a year. The bird keeps a long tail through the period of molt. Tail feathers are replaced with one of the three long rectrices T1-2-3 which are retained until another long feather on each side is completely replaced. In some coucals (best known in Pheasant Coucal Centropus phasianinus in Australia) the central rectrices T1 are replaced twice and the others once a year (Higgins 1999). Smooth-billed Anis Crotophaga ani have a complete post-juvenile body and tail molt and a partial wing molt that is similar in pattern and extent to the adult molt. In March and April the young birds undergo a partial molt, but they attain the wide adult tail feathers only after the second autumn molt (Dickey and Van Rossem

52 The Cuckoos 1938). Adult anis undergo a complete postbreeding molt and a partial prebreeding body molt, with the molt slow and lasting through much of the year, and birds sometimes breed while they are still in molt (Davis 1940a, Snow and Snow 1964, Foster 1975, Pyle 1997). Molt is generally seasonal and occurs when birds are not breeding or in migration. Molt is not a reliable guide to the seasonality of breeding in cuckoos, and some cuckoos molt while they are breeding. In addition to the anis, Lesser Ground-cuckoo Morococcyx erythropygus and Dwarf Cuckoo Coccycua pumila sometimes molt while they are in breeding condition (Foster 1975, Ralph 1975), breeding adult Yellow-billed Cuckoos sometimes begin postbreeding molt while they are still caring for their nestlings (Potter 1980), and a few brood-parasitic females in Africa begin to molt while still laying (Payne 1969a). Molt in cuckoos is a prolonged affair. Migratory cuckoos such as Common Cuckoo begin and complete their post-juvenile and post-breeding molts on their wintering grounds. In northern temperate-region cuckoos that migrate, the molt is sometimes incomplete or is delayed for months. Common Cuckoos replace their plumage in about 100 days (Seel 1984b). For the Diederik Cuckoo and Klaas’s Cuckoo, the juvenile primary molt begins at 4 months and is completed in about 80 days (Hanmer 1995). In several species including the Common Koel, Diederik Cuckoo and Klaas’s Cuckoo, juvenile molt is often interrupted with birds in their first year ending the molt with their wing primaries, secondaries and coverts a mix of juvenile and adult feathers. Most remarkable in cuckoos is the molt pattern in which the flight feathers are replaced. The wing feathers are replaced not from the innermost outward, but from separate centers within the wing, beginning with an outer primary (Stresemann and Stresemann 1961, 1966, 1969, Piechocki 1971, Stephan 1991). Molt proceeds in leaps forwards or backwards across one or more adjacent primaries, in a “transilient” pattern. As with other birds, the primaries are numbered from the innermost feather outward; the terms “ascending” and “descending” trace back to an earlier time when the feathers were

numbered from the outermost feather inward. The Old World parasitic cuckoos shed the primaries in a predictable sequence. In the simple form of transilient molt, the odd-numbered feathers are replaced first from outer to inner position, then the evennumbered feathers are replaced, again starting from the outer position. Several variations occur on this theme: in Cuculus, Cacomantis and Chrysococcyx the molt sequence is P9 - 7 - 5 - 8 - 6, with P 4 - 1 2 - 3 beginning between the old primaries P5 and P3. Long-tailed Cuckoo Urodynamis taitensis has a unique molt, first the odd-numbered primaries then the even-numbered primaries, with the molt sequence P9 - 7 - 5 - 3 - 10 - 8 - 6 - 4. Certain cuckoos have other patterns, with Common Koel Eudynamys scolopacea quite different from other parasitic cuckoos with the outer primaries in transilient ascending molt (P9 - 7 - 5 - 10 - 8 - 6) and the inner primaries in stepwise descending molt (1 - 2 - 3 - 4) (Stresemann and Stresemann 1961). Another pattern is a semi-transilient ascending molt where some adjacent feathers are molted (Scythrops,Tapera). The crested cuckoos Clamator and lizard-cuckoos Coccyzus (Saurothera) are unique as their wing molt progresses in a regular pattern to skip two feathers, with a sequence P6 - 9 - pause - 7 - 10 - pause - 8 - 5 (or - 5 - 8) (Stresemann and Stresemann 1969, Piechocki 1971). In the Great Spotted Cuckoo these feathers are of different colors in juvenile and adult plumage, and first-year birds can be identified by their primaries, the new odd-numbered adult primaries being gray-brown and the evennumbered primaries the old rufous juvenile feathers. Non-parasitic cuckoos have a variable pattern of transilient ascending molt where adjacent feathers are sometimes replaced in series (Coccyzus, Ceuthmochares, Centropus) but the same pattern of alternating juvenile and adult feathers is sometimes seen in the flight feathers of first-year birds (Pyle 1997). Molt sequences sometimes vary within a species (Stresemann and Stresemann 1961) and the number of museum specimens in active molt is limited, so it is not possible at this time to use molt patterns as characters in a phylogenetic analysis for all cuckoos. This unique transilient pattern of molt nevertheless supports the monophyly of cuckoos.

Morphology 53

Skeleton Avian anatomists and paleontologists identify birds from the details of their bones (Huxley 1867, Fürbinger 1888, Pycraft 1901, 1903, Howard 1929, Berger 1960, Olson 1985, 1990, Olson and James 1991, Feduccia 1996). Certain bones have characteristic shapes that differ among avian families and orders, and skeletons may be useful in discovering the evolutionary relationships among birds of the world. It is only since the application of cladistic reasoning or phylogenetic systematics (Hennig 1966) that ornithologists have distinguished between characters that are shared because they are primitive (inherited from a remote ancestor, and shared only because other lineages from the same ancestor have evolved different characters, and not because they are shared by a strictly monophyletic lineage), or because they are derived (synapomorphic, inherited from a unique and recent common ancestor), and that skeletal characters have been used in the logic of phylogenetic systematics (e.g. Cracraft 1974, Payne and Risley 1976, Livezey 1993, 1997, 1998, Mayr 1998a,b). Despite the accumulation of a vast amount of descriptive detail on skeletal features, systematists have only recently begun to organize the accumulated data in a phylogenetic research program to determine the evolutionary relationships among birds (Livezey and Zusi 2001). In practice, bird skeletons are identified by comparison with a reference collection where bones are matched for identification to samples without a detailed description, and the published descriptions of bones are usually limited to certain groups or to a geographic region. While these resources are helpful, they often do not include all variations found within a family, and identification is sometimes based on a process of elimination (“not as above”) (e.g. Gilbert et al. 1981). Cuckoos have a desmognathous palate, no vomer, no basipterygoid process, a holorhinal nasal septum, deep temporal fossae and short mandibular processes.The postcranial skeleton has 13 or 14 cervical vertebrae.The pectoral girdle has a large coracoid and the sternum has an internal spine.The pubis has a pectineal or preacetabular process.The leg has two bony canals arranged side by side in the hypotarsus (the crest complex on the proximal end of the

tarsometatarsus), and the foot is zygodactyl, with two toes forward and the inner and outer toe directed backward, and with an accessory process on trochlea IV of the distal end of the tarsometatarsus to support the backwards orientation of the outer toe. These characters of cuckoos have been pointed out in earlier studies as reviewed by Sibley and Ahlquist (1990), and they were observed in skeletons in the present study. These skeletal features taken one at a time are not unique to cuckoos. For example, a desmognathous palate (a palate with a medial bridge across the maxillopalatinos: Huxley 1867, Pycraft 1901) also occurs in ducks, falcons, turacos, parrots and rollers and a few other avian groups, so this palate structure has evolved more than once in birds (Huxley 1867, Pycraft 1901). Other skeletal features vary among the cuckoos, such as the pectineal process of the synsacrum, a process which is prominent only in the terrestrial cuckoos (Verheyen 1956).There has been little consideration of whether the skeletal features of cuckoos are uniquely derived (synapomorphies) within the group, or are shared with other avian groups with a common ancestry, or are independently derived in the cuckoos and in other groups (homoplasies). Characters of cuckoos that may have been independently derived in bird lineages include the zygodactyl foot, a foot structure that also occurs in parrots, owls and woodpeckers. Although these birds may not form a monophyletic lineage, they are related in one phylogenetic hypothesis (Sibley and Ahlquist 1990), and if that model is true, then zygodactyly was ancestral to the original land-bird radiation and was subsequently lost three times (in Coraciiformes; in nightjars Caprimulgiformes, some swifts and hummingbirds Apodiformes; and in the lineage leading to pigeons, rails, the wading birds Ciconiiformes and the perching birds Passeriformes). The details of the zygodactyl foot in cuckoos differ from those in other zygodactyl birds and other birds that have two toes directed forwards and two backwards (trogons, mousebirds, some swifts, osprey, parrots, owls, cuckoo-roller, piciforms and the jacamars and puffbirds) (Figure 4.11). In zygodactyl birds the outer and inner toes (digits I and IV) are directed backwards; in fact the outer toe of

54 The Cuckoos

Figure 4.11. Distal end of the tarsometatarsus of cuckoos, hoatzin, turaco and other zygodactyl birds: a, cuckoo Geococcyx californianus; b, hoatzin Opisthocomus hoazin; c, turaco Tauraco hartlaubi; d, parrot Amazona amazonica; e, owl Otus asio; f, woodpecker Colaptes auratus. ii ⫽ trochlea II; iii ⫽ trochlea III; iv ⫽ trochlea IV; Se ⫽ sehnenhalter. Each pair of figures shows the end view (above) and ventral view (below), left tarsometatarsus; scale mark is 3 mm.

climbing woodpeckers is also often directed laterally (Bock and Miller 1959). The distal end of the tarsometatarsus in these zygodactyl birds with digit IV permanently reversed (cuckoos, parrots, owls, piciforms, jacamars and puffbirds) has an accessory articulating process for digit IV, a large inflected collateral flange, the sehnenhalter or sinew holder (Steinbacher 1935, Simpson and Cracraft 1981, Olson 1983, Houde and Olson 1992), and in mousebirds the simple outer flange of trochlea IV is enlarged and birds can reverse the outer toe even

though the toe is not permanently reversed (Raikow 1985).This sehnenhalter process in zygodactyl birds is enlarged and rotated posteriomedially. In cuckoos the process is less distinct than in other zygodactyl birds. It extends posteriomedially from trochlea IV but not as far distally as trochlea IV (it extends further distally in parrots), its lateral length is less than half the width of the distal shaft of the tarsometatarsus (it is about half the width in jacamars and puffbirds (not shown), and is longer in woodpeckers), and it projects on the perpendicular

Morphology 55 only slightly from the shaft of the tarsometatarsus (it projects radially a distance more than the width of the tarsometatarsus in woodpeckers and other piciforms, in which it also extends along the shaft further than its radial projection). The structure is similar for all cuckoos though it is slightly longer in the terrestrial cuckoos such as roadrunner. In hoatzin, trochlea IV bears a posteriolateral process which is not however an extra process (not a sehnenhalter), and it is similar to the expansion of the lateral ridge of trochlea IV in the trumpeter (Psophia crepitans UMMZ 152877, Psophiidae) which has an even weaker sulcus for the trochlear IV tendon; the hoatzin is not zygodactyl and the outer three toes are directed forwards.Turacos have a semi-zygodactyl foot in which the outer toe can go forwards, it is usually directed outwards, and it can rotate backwards within 70o of the hallux (Moreau 1938); this flexibility of the outer toe does not involve a sehnenhalter (Fig. 6). In the heterodactyl trogons digits I and II are sometimes directed backwards, not digits I and IV as in the zygodactyl birds; actually trogons usually hold all four toes forwards. The arrangement of foot tendons co-varies with the tarsometatarsal trochleae in zygodactyl birds (Gadow 1892, Steinbacher 1935, Olson 1983), except that muscle attachments are similar in jacamars and puffbirds and in the typical piciform birds (Swierczewski and Raikow 1981, Raikow and Cracraft 1983). The mechanical function of this process is to articulate with the basal phalanx of digit IV in the reversed position and to reverse the direction of force of the long flexor tendons of digit IV (Scharnke 1930, Steinbacher 1935). The foot of the cuckoo is also distinctive at the proximal end of the tarsometatarsus, where tendons from the leg to the toes pass through two bony canals in the hypotarsus.The two canals are arranged side by side in cuckoos. Other zygodactyl birds have a differently structured hypotarsus. In woodpeckers two bony canals are arranged radially, owls have no bony canals, and in parrots the hypotarsus varies with a single canal in Aratinga and Ara, a single canal with a partial division into two radial canals in lories Lorius garrulus and two canals side by side with incomplete division of one of them in the cockatoo

Cacatua moluccensis. In addition a few other birds including a grebe Podiceps grisegena and some herons Botaurus (Payne and Risley 1976) have two bony canals side by side, and songbirds have a hypotarsus with two canals side by side and also have other canals arranged radially to these (Figure 4.12). Although the foot skeleton of cuckoos is distinctive, the presence of two bony canals that are arranged side by side in the hypotarsus is not unique. Cuckoos have 13 or 14 cervical vertebrae and the number has figured prominently in earlier systematic arrangements of cuckoos. Fürbinger (1888) reported 13 in Crotophaga and 14 in Cuculus, Zanclostomus javanicus and Centropus; Shufeldt (1886a,b, 1901) and Berger (1952, 1960) reported 13 in Clamator, Coccyzus (including Saurothera) and Piaya, and 14 in most other cuckoos; Hughes (2000) reported 13 only in Clamator and Coccyzus. Pycraft (1903) recognized only 11 or 12 but he also recognized 2 or 3 cervico-thoracic vertebrae; Newton (1896) and Gadow and Selenka (1893) recognized 14 or 15 cervical vertebrae in cuckoos. The number of cervical vertebrae depends on the criteria used to define them. As Beddard (1899a) noted, the last three neck vertebrae bear ribs that progressively increase in size, with the anteriormost ones being very small.The ribs articulate anteriorly with the head, and a facet on the short knob on the anteriolateral surface of the vertebral centrum, and posteriorly with the tubercle and a larger process of the neural arch. Fürbinger (1888) and Baumel (1993), distinguished cervical vertebrae from thoracic vertebrae in that they do not articulate with a complete rib, where a “complete rib” has both vertebral and sternal segments, and the sternal segment connects directly or indirectly with the sternum. Vertebrae at the root of the neck which bear moveable ribs that do not reach the sternum are transitional between cervical and thoracic vertebrae. Newton (1896) and Berger (1960) considered only vertebrae with a rib that have a direct connection with the sternum to be thoracic vertebrae, yet they found different numbers of cervical vertebrae. Pycraft (1903) was more restrictive. He considered the vertebrae behind the cervicals to be cervicothoracic vertebrae, recognized by their ventral catapophyses in addition to the median hypapophysis,

56 The Cuckoos

Figure 4.12. Proximal end of the tarsometatarsus, showing the canals in the hypotarsus: a, cuckoo Geococcyx californianus; b, hoatzin Opisthocomus hoazin; c, turaco Tauraco hartlaubi; d, songbird Corvus brachyrhynchos; e, grebe Podiceps grisegena; f, woodpecker Colaptes auratus; g, owl Otus asio. c ⫽ canals; h ⫽ hypotarsus. Each pair of figures shows the end view (above) and ventral view (below), left tarsometatarsus; scale mark is 3 mm.

the second and third cervico-thoracics having ribs and the third having a rib with an uncinate process. He also noted that the fifth or posteriormost thoracic rib did not reach the sternum.Another feature that has been used to differentiate the cervical and thoracic vertebrae is the shape of the dorsal spine, which is short in the axial plane in the cervical vertebrae and has a high longitudinal crest in the dorsal vertebrae (Newton 1896), whereas Pycraft (1903) noted that the dorsal spine of the first thoracic vertebra in Coua is barely visible and is like that of the posterior cervical vertebrae. Using Fürbinger’s definition, in skeletal specimens where the articulation of ribs from the vertebrae to the sternum is intact I recognize 13 cervical vertebrae in Surniculus and Cuculus as well as in Clamator, Zanclostomus javanicus and Carpococcyx renauldi. Inasmuch as (a) some studies did not describe the criteria to recognize the cervical vertebrae, (b) the

transitional series of these vertebrae are associated with the transitional series of the ribs, (c) the attachment of the rib to the sternum is not obvious in disarticulated specimens, and (d) the number of cervical vertebrae sometimes varies among specimens of a single species both in cuckoos and in other birds (Fürbinger 1888, Berger 1952, 1960), the definitions and homologies of avian cervical vertebrae need to be re-evaluated. The sternum varies in the depth of the keel, where the flight muscles attach. The sternum also varies in the shape of the posterior edge (notched, windowed or intact).The keel is short and shallow in terrestrial ground cuckoos such as the couas and coucals (Pycraft 1903). These cuckoos can fly. The flight muscles attach not only to the keel but also to a stout coraco-clavicular membrane that extends from the coracoid to the furcula, as in Pheasant Coucal (Feduccia 1996, p. 136).

Morphology 57

Figure 4.13. Humerus of cuckoos and other birds: a, Geococcyx californianus; b, Eudynamys scolopacea; c, Opisthocomus hoazin; d, Tauraco hartlaubi; e, Trogon citreolus; f, Colaptes auratus; g, Indicator indicator; h, Corvus brachyrhynchos. aet ⫽ anterior external tuberosity; de ⫽ deltoid crest; ect ⫽ ectepicondyle; ent ⫽ entepicondyle; head ⫽ head; rt ⫽ round tuberosity; s ⫽ shaft. Anconal view, right humerus; scale mark is 5 mm.

In cuckoos the humerus is concave along the medial edge; it has a strongly developed median crest, and it lacks the distinct round tuberosity proximal to the entepicondyle that characterizes the perching birds Passeriformes. The humerus is distinct in the shape of the deltoid crest.The crest projects in a ridge with two peaks, a minor peak located proximally at the anterior external tuberosity and a larger peak located distally on the crest with a shallow concave area between the peaks (Figure 4.13). The crest is short and does not extend far along the shaft, has a simple low ridge and often a more extensive distal projection along the shaft.The distal peak is more prominent in Eudynamys and less

prominent in the ground-cuckoos Carpococcyx and Neomorphus. In other groups of birds, the deltoid crest is a straight ridge with a single peak (songbird), or it has a single peak without the extensive ridge ( parrot, pigeon), or a single rounded ridge (mousebird, not shown in Figure 4.13), while other details in the humerus also differ between cuckoos and other birds.

Skeletons and cuckoo monophyly Seibel (1988) and Hughes (2000) described several features of cuckoo skeletons that they interpreted as unique to the cuckoos, and other features that

58 The Cuckoos characterized certain lineages, including terrestrial vs arboreal cuckoos, and the brood parasites. These features involved the shape and combination of skeletal elements, and not the presence or absence of a simple element. Hughes used the more inclusive set: 14 characters. Because these skeletal characters were used to estimate the evolutionary history of cuckoos, and the conclusions of these studies differed from conclusions of studies in molecular genetics, it is of interest to compare these characters. The skeletons of cuckoos and other avian taxa were compared, in the present study, to test whether the characters reported to support monophyly of cuckoos were in fact unique to cuckoos. One or two representatives from each major cuckoo clade were selected. The other taxa were selected from specimens in orders that were within three branches of the cuckoos in the phylogenetic estimate of Sibley and Ahlquist (1990): mousebirds Coliiformes, kingfishers Coraciiformes, puffbirds Bucconidae, buttonquail Turniciformes, woodpeckers and honeyguides Piciformes, parrots Psittaciformes, turacos Musophagiformes, owls Strigiformes, doves Columbiformes, plovers Charadriiformes, falcons Falconiformes and herons Ciconiiformes (these last three all classed as Ciconiiformes in Sibley and Ahlquist 1990) and songbirds Passeriformes. The sample was selected to include at least one taxon from all third-order branches (clades within three branching points) of the cuckoos in Sibley and Ahlquist (1990), and from each order that has been suggested elsewhere as being closely related to the cuckoos.The sample also included hoatzin, which was regarded by Sibley and Ahlquist as a cuckoo.When a bone was missing or damaged, another specimen was examined. Taxa and specimens examined are listed in Table 4.3, and the characters of cuckoo monophyly in Seibel (1988) and Hughes (2000) are described in Table 4.4. The examination of skeletons does not support the claim (Hughes 1990) that all 14 characters are unique to cuckoos (Table 4.4). Depending on the character, the structures vary among the cuckoos, or they are not unique to cuckoos, or they vary both among cuckoos and among other taxa, although a few characters are unique. Cuckoos and the other

taxa differed in (8): the occurrence and position of two hypotarsal canals side by side.A few other birds also have two hypotarsal canals arranged side by side, so this characteristic is not unique to cuckoos. The sehnenhalter or accessory trochlea IV differs between cuckoos and other zygodactyl birds, but the description of character (9) does not make this distinction. The shape of the maxillary bone is distinctive, as is the form of the humerus. The other characters are variable among cuckoos or also occur in other taxa.

Skeletons and evolutionary relationships between cuckoo species Comparison of cuckoo skeletons has led to estimates of cuckoo relationships that affect the interpretation of important questions in evolution, including the origin of brood parasitism, the adaptations of arboreal and ground cuckoos, and the interchange between or isolation of the Old and New World cuckoos (Seibel 1988, Hughes 1997d, 2000). This phylogenetic inference is based on the assumption that truth is revealed in bones, and skeletal characters are true guides to the evolutionary relationships within the cuckoos.

Single origin of arboreal and terrestrial behavior? In an analysis of the postcranial skeleton of cuckoos, Seibel (1988) concluded that ground-cuckoos of the Old World were closely related to groundcuckoos of the New World, and neither were closely related to the arboreal cuckoos in their regions. The characters that linked these groundcuckoos were features of the synsacrum and tarsometatarsus, and these may have independently derived their structure in relation to the terrestrial mode of locomotion. Seibel reported three traits as shared derived characters for the recognition of a clade of coucals, couas and the Old World and New World ground-cuckoos: the relative length of the ilioischiatic synsacrum (long in these cuckoos), and the contour and orientation of the anterior edge of the posteriodorsal iliac crest (straight and more

Morphology 59 Table 4.3 Skeletal specimens examined to test statements that cuckoos are diagnostically distinct. Cuckoos Carpococcyx renauldi Centropus senegalensis Centropus toulou Clamator glandarius Coccyzus americanus Coua caerulea Crotophaga ani Cuculus canorus Dasylophus superciliosus Eudynamys scolopacea Geococcyx californianus Neomorphus geoffroyi Zanclostomus javanicus Surniculus velutinus Other taxa Amazona amazonica (parrot) Ardea herodias (heron) Ceryle alcyon (kingfisher) Charadrius vociferus (plover) Colaptes auratus (woodpecker) Colius striatus (mousebird) Columba livia (pigeon) Corvus brachyrhynchos (songbird) Falco sparverius (falcon) Indicator indicator (honeyguide) Notharchus macrorhynchos (puffbird) Opisthocomus hoazin (hoatzin) Otus asio (owl) Tauraco hartlaubi (turaco) Trogon citreolus (trogon) Turnix suscitator (buttonquail)

laterally directed in these cuckoos). This first character is problematic, as most cuckoos had an intermediate form (Seibel 1988). Second, the couas, Old World ground-cuckoos and New World groundcuckoos were uniquely linked only by the inflection of the acrocoracoidial furcular facet of the coracoid, although the definition of character states of the facet is problematic (Seibel 1988). Third,

specimen number UMMZ 219851, 236029 UMMZ 207709 UMMZ 208402 UMMZ 212907 UMMZ 152712 UMMZ 209201, 208404 UMMZ 218940 UMMZ 151115 UMMZ 228022 UMMZ 210347 UMMZ 71909 UMMZ 200592 UMMZ 236439, 236440 UMMZ 233061 UMMZ 203942, 203944 UMMZ 107419, 224551 UMMZ 206751 UMMZ 74184, 198342, 220113 UMMZ 235110, 207776 UMMZ 234219 UMMZ 224018 UMMZ 232727 UMMZ 223964 UMMZ 212931 UMMZ 219195 LSU 62741 UMMZ 224665 UMMZ 234273 UMMZ 209256 UMMZ 205093

within this last set of cuckoos, the Old World ground-cuckoos and New World ground-cuckoos including roadrunners were linked by two tarsometatarsal traits: the size and position of a tubercle, and the position of the distal edge of the proximal foramen. Insofar as the characters are in the synsacrum and legs (or are problematic), and these birds share a common pattern of locomotion,

60 The Cuckoos Table 4.4 Skeletal characters in a statement of cuckoo monophyly (after Hughes 2000). 1. “Os quadratum. Condylus medialis, caudalis, and lateralis prominent, well-rounded, and separated by a broad notch; condylus caudalis tapered to a rounded point, deflected and extending well beyond caudal margins of quadrate in lateral aspect; processus orbitalis quadrati truncated and somewhat anvil-shaped.” (Character 21 of Hughes 2000). Comments:The description is appropriate for cuckoos, although the shape of the condyles vary considerably in detail, particularly the shape of the orbital process, and the condylus medialis is not demarcate by a distinct notch in relation to the other quadrate condyles. In the other taxa, the same description appears to match the bones in part, if not in entirety: (1a) the condyles are prominent and well rounded in the parrot, kingfisher, trogon, heron and songbird; (1b) the broad notch is present in the buttonquail, trogon, kingfisher, mousebird, hoatzin, owl, turaco and songbird, and the heron has a prominent but narrow notch; (1c) the condylus caudalis is tapered to a rounded point and deflected in the kingfisher, plover, songbird, owl and turaco; (1d) the condylus caudalis extends as in cuckoos, in the honeyguide, kingfisher, trogon, turaco, owl, plover and songbird; (1e) the processus orbitalis quadrati is truncated in the buttonquail, hoatzin, turaco, parrot, plover, falcon and songbird; and (1f ) the processus orbitalis is anvil- or cone-shaped in the hoatzin, parrot, plover, dove and falcon. Because these characters are variable within the cuckoo species, and none are unique to cuckoos, the form of the quadrate is not diagnostic in cuckoos, i.e. it does not clearly distinguish them. 2. “Os pterygoideum. Mesially deflecting spur-like processus dorsalis located caudally on facies dorsalis near, but not articulating with, os quadratum.” (Character 44 of Hughes 2000). Comments: Not all cuckoos have such a process, and in Centropus, Clamator and Eudynamys the process articulates with the quadrate. 3. “Os ectethmoidale. Large, quadrilateral in form, marginis lateralis and ventralis slightly to moderately excised.” (Character 15 of Hughes 2000). Comments:The character has three components: in cuckoos (3a) all have a quadrilateral ectethmoid; (3b) a lateral excision is apparent only in Clamator, Coccyzus, Eudynamys, Geococcyx and Neomorphus; and (3c) a ventral excision is apparent in all but Carpococcyx, Centropus, Coua and Dasylophus. In other taxa, seven have a quadrilateral ectethmoid, five have a lateral excision, and six have a ventral excision; the hoatzin is deeply excised (not slightly to moderately excised), the puffbird is similar to Centropus, although the ectethmoid is extended further laterally, the mousebird is nearly identical to Coccyzus, and the buttonquail is nearly identical to Coua. 4. “Os palatinum. Marginis medialis meeting in midline below, fully or nearly concealing, rostrum parasphenoidale; lamella caudolateralis being widest midway between processus pterygoideus and margo distalis of pars choanalis, tapering towards rostrum; processus interpalatinus extending nearly to margo caudalis of processus maxillopalatinus; in lateral aspect, cristae lateralis are deflected ventrally.” (Character 47 of Hughes 2000). Comments:The description defines the desmognathous palate and elaborates on the shape of the palatines. In cuckoos the description is appropriate, except that (4b) the parasphenoid is exposed throughout its length in Carpococcyx; and (4d) the palatines do not taper anteriorly in Clamator.The description excludes the other taxa, as most of these including hoatzin are schizognathous, although in all (4d) the palatines taper to the rostrum and (4f) the lateral crests are deflected ventrally in all but kingfisher. 5. “Os maxillare. In ventral aspect, processus palatinus bilobed, extending caudally nearly to marginis distalis of crista ventralis; margo medialis of processus palatinus not fused caudally to margo distalis of os maxillare.” (Character 48 of Hughes 2000). Comments: The description applies to cuckoos. In the other taxa, it also applies to plover, songbird and buttonquail, although the shape of the maxilla differs in other aspects from that of cuckoos. 6. “Os articulare, crista intercotylaris. Prominent, half-disk shaped with radius aligned approximately 45° to planum medianum.” (Character 54 of Hughes 2000). Comments:The description of the base of the lower mandible applies to cuckoos. In other taxa, the description does not universally apply as the articular surface is more complex (more robust in parrot) and the radius is closer to the median plane in turaco, owl, falcon and buttonquail, but the description applies well to the hoatzin, puffbird and trogon which are nearly identical to cuckoos. 7. “Tibiotarsus, incisura intercondylaris. Deep, rounded pit extending caudally at least half the length of extremitas distalis; trochlea cartilaginis tibialis poorly developed.” (Character 57 of Hughes 2000).

Morphology 61 Table 4.4 contd. Comments: This character has two components: (7a) the extent of the pit; and (7b) the development of the trochlea. The description of the distal end of the tibiotarsus applies to cuckoos. In the other taxa both 7a and 7b apply to the mousebird, turaco and owl. 8. “Tarsometatarsus, hypotarsi. Two oblong canales hypotarsi, completely enclosed in bone, positioned side by side in planum medianum both almost entirely caudal to cotylae medialis and lateralis, and corpus tarsometatarsi.” (Character 58 of Hughes 2000). Comments:This character is present in the cuckoos, while in the other taxa examined it is absent. Some have a single canal, e.g. hoatzin. However certain other birds have two enclosed hypotarsal canals, either side by side (some grebes) or in another arrangement (not side by side: woodpeckers, some herons, Payne and Risley 1976), or have two canals in the same position as the cuckoos, but have additional canals peripheral to those in a radially expanded hypotarsus (songbird) (Fig. 5). 9. “Tarsometatarsus, trochlea metatarsi quarti. Margo distalis proximal to incisura intertrochlearis medialis; caudal trochlea accessoria prominent and strongly inflected medially.” (Character 75 of Hughes 2000). Comments:This character has two components: (9a) the shape of the distal margin; and (9b) the sehnenhalter (accessory caudal trochlea). All cuckoos have the character, while in the other taxa, (9b) the sehnenhalter is inflected inwards in the owl, woodpecker and honeyguide, as in cuckoos (in owl the outer trochlea IV flange forms the base of the structure, and in woodpecker and honeyguide the inner trochlea IV flange forms this structure as in cuckoos; in parrots trochlea IV is shorter—it does not extend very far distally along the shaft—than the sehnenhalter). In hoatzin the outer trochlear flange is prominent, but less so than in cuckoos (in hoatzin, trochlea IV extends as far distally as trochlea II, the flange is about 60% as long as the metatarsal articular surface of trochlea IV, and the flange not inflected inwards (Fig. 4). 10. “Coradoiceum, facies articularis sternalis. Crista ventralis prominent, centred, rounded, and facing sternally.” (Character 99 of Hughes 2000). Comments: Gilbert et al. (1981) describe this condition as “ventral sternal facet rounded, centered and facing sternally”. Although the figures in Gilbert et al. (1981) and Hughes (2000) do not show the feature well, the coracoid has a ridge or crest on the ventral surface, perpendicular to the shaft and centered with respect to the shaft, with the curved ridge parallel to that of the margin of the sternal articulatory facet. All cuckoos have a coracoid as in this description, while in the other taxa, trogon and plover have the crest as in cuckoos, the woodpecker is similar (though the crest is more distal to the coracoid shaft), and the pigeon has a prominent ventral crest not centered but expanded on the procoracoideal side of the coracoid. The ventral crest is also present in certain other birds including herons and cormorants (Phalacrocorax auritus) although it is not always centered on the shaft in these birds (Payne and Risley 1976). It is also known in other birds including early fossil charadriiform, trogon and other Eocene fossils (Mayr 1998a, 1999, 2000). In hoatzin the coracoid is fused with the sternum and no articular surface is visible. Seibel (1988) recognized fusion of the bones and assumed that hoatzin lacked a ventral crest (his character state CO19a); Hughes (2000) did not describe hoatzin though she coded the crest as absent in hoatzin. 11. “Scapula. Facies articularis humeralis directed dorsally; facies articularis clavicularis prominent, knob-like, and directed dorsally; extremitas cranialis scapulae, delineated by acromion and tuberculum coracoideum, truncated and flattened cranially.” (Character 107 of Hughes 2000). Comments:The description fits the cuckoos, and also the other taxa, except that (11a) the facies articularis humeralis is directed ventrally in the kingfisher; (11b) the facies articularis clavicularis is not prominent in the songbird, and (11c) the facies articularis clavicularis is not directed dorsally in the plover, while in the woodpecker the caudal end of the scapula is greatly enlarged, flattened and recurved to meet the rib cage. 12. “Humerus, condylus ventralis. Rounded, not flattened or oblong, in axis proximodistalis and both plana transversialia and dorsalia.” (Character 114 of Hughes 2000). Comments: In cuckoos the description matches the specimens, while in the other taxa it matches for all except woodpecker, honeyguide and trogon. 13. “Synsacrum, extremitas caudalis synsacri. Crista dorsalis reduced and fused to form a single low ridge; marginis lateralis of processus transversus widened in axis rostrocaudalis, facies lateralis projecting cranially.” (Character 130 of Hughes 2000).

62 The Cuckoos Table 4.4 contd. Comments: In cuckoos, (13a) the crista is not prominent in Crotophaga; and (13d) there is no wide lateral margin of the transverse process in Clamator and the lateral margin does not project cranially in Neomorphus, while in the other taxa, (13a) the crista dorsalis is reduced in the kingfisher, songbird, owl, turaco and buttonquail; (13b) the crista dorsalis is unfused in all but the plover; (13c) a single low ridge is present in all but the songbird and hoatzin; (13d) there is a wide lateral margin of the transverse process in the mousebird and honeyguide; and (13e) the lateral face projects cranially in the plover and turaco. 14. “Vertebrae caudales, primus. Processes transversus tapered, not widened, paddle-shaped in ventral aspect; projecting strongly caudally.” (Character 131 of Hughes 2000). Comments: The description is appropriate for most cuckoos, although (14a) the transverse process is not tapered in Carpococcyx, Coua, Dasylophus or Eudynamys; (14b) it is not paddle-shaped in Coua; and (14c) it does not project strongly caudally in Carpococcyx, Coua or Eudynamys, while in the other taxa, (14a) the process is tapered in the kingfisher, trogon, songbird, falcon and turaco; (14b) it is paddle shaped in the woodpecker, mousebird and hoatzin; and (14c) it projects strongly caudally in the parrot, mousebird and owl.

the skeletal characters are likely to be adaptations for running on the ground, and evolved independently in these cuckoos (Payne 1997b). Only one character was found to link the arboreal cuckoos: the posterior extent of the medial angle of the posterior dorsal iliac crest, relative to the width of the vertebral centra (the main body of the vertebra).The cuckoos that share the character are the New World anis, New World broodparasitic cuckoos, New World coccyzine cuckoos and the Old World malkohas and brood-parasitic cuckoos. Seibel’s conclusion that a synsacral character is a derived synapomorphy may follow from the assumption of turaco as the outgroup (Payne 1997b). These long-legged birds have synsacra that are used in bounding locomotion, so it is questionable whether this trait is a good standard by which to estimate the evolutionary polarity of characters in cuckoos. The shape of the iliac crest in these cuckoos may be a primitive rather than a derived character. Also, it is uncertain that turacos are closely related to cuckoos.With only a single skeletal character to link these cuckoos with each other, there is no corroboration of the hypothesis that arboreal cuckoos form a single lineage.

Single origin of brood parasitism? Skeletal characters were used by Seibel (1988) and Hughes (2000) to infer a single origin of broodparasitic cuckoos.These characters were re-examined

in the present study, using some of the same specimens and others that were not available to the earlier studies. Seibel (1988) listed four characters that defined the brood-parasitic cuckoos of the Old World and New World as a single derived group, together with the New World Coccyzus cuckoos, and two additional characters that defined this group except for the parasitic genera Urodynamis and Eudynamys. Hughes (2000) used a larger set of 135 skeletal characters, including 56 cranial as well as 79 postcranial characters, and she modified several characters of Seibel (1988). In addition some taxa were different (Dasylophus cumingi in Hughes 2000, Zanclostomus javanicus in Seibel 1988). The phylogenetic estimates of the two studies were similar, with a few exceptions. In Hughes’ estimate, (1) Carpococcyx was basal to all cuckoos, (2) Coua and Centropus formed the next basal clade, and (3) the New World Morococcyx, Geococcyx and Neomorphus formed the next basal level (rather than a sister clade to Old World Carpococcyx).The differences in (1), (2) and (3) may result from rooting the outgroup next to Carpococcyx rather than between the ground cuckoos and arboreal cuckoos. Other differences in estimates were that in Hughes (2000), (4) the mostly Old World Phaenicophaeinae and Cuculinae formed separate clades, and (5) New World brood-parasitic Tapera and Dromococcyx were basal to the Old World brood-parasites with both in Cuculinae. Hughes’ estimate was similar to Seibel’s in (1) the association of New World Coccyzus with

Morphology 63 the Old World brood-parasitic cuckoos rather than with New World nesting Piaya, and (Coccyzus sub-genera) Saurothera and Hyetornis, and (2) the association of Morococcyx, Geococcyx and Neomorphus in the New World ground cuckoos. Hughes (2000) included a data matrix, and this matrix and a character analysis were presented elsewhere as an appendix (Hughes 1997d ). Both Seibel (1988) and Hughes (2000) concluded that brood parasitism evolved only once in the cuckoos. Hughes’ presumptive clade of “Cuculinae” (Old World and New World brood parasites together with the genus Coccyzus) was distinguished in Hughes (1997d, 2000) by nine skeletal characters that changed state from node #63 to node #64, and defined her “Cuculinae” parasitic cuckoos. To evaluate this estimation of a single origin of brood-parasitic cuckoos, I examined skeletons of cuckoos for the six characters that were said to define the brood-parasitic clade in Seibel (1988), and the nine characters said to define the clade in Hughes (1997d, 2000), a total of 12 characters, with some overlap between the two studies. The terminology and examples refer to the illustrations and descriptions in standard anatomical works (Howard 1929, Baumel 1993).

Cuckoo skeletons were examined to test their conformity with coding of the characters that Seibel and Hughes reported, to support their conclusion of a single origin of brood-parasitic cuckoos. Specimens used for the examination were: Clamator glandarius 212909, Cuculus canorus 2020512, 119430, Eudynamys scolopacea UMMZ 210347, Urodynamis taitensis USNM 559589, Tapera naevia UMMZ 214005, Dromococcyx pavoninus UMMZ 209207, Crotophaga major UMMZ 219553, Geococcyx californianus 227057, Morococcyx erythropygus UMMZ 133738, Ceuthmochares australis UMMZ 158186, Zanclostomus javanicus UMMZ 209478, Dasylophus superciliosus UMMZ 228024, Coccyzus americanus UMMZ 154465, Saurothera merlini UMMZ 158527, Coccycua minuta UMMZ 139991, 139990, Piaya cayana UMMZ 200591, Piaya melanogaster UMMZ 209478, Coua caerulea 209201, Coua cristata UMMZ 157526, and Centropus senegalensis UMMZ 205896. Where skeletal elements were missing, a second specimen was examined. In most cases there was no consistent or discrete difference in the skeletal trait for nesting cuckoos and brood-parasitic cuckoos, as in Table 4.5. Where the CI is less than 1.0, the character transition occurred more than once in

Table 4.5 Skeletal characters in a statement of monophyly of brood parasitism (“Cu”) in cuckoos (Hughes 1997d (node 66), 2000). Characters are numbered as in Hughes (1997d ), where the character transitions are described. Character number, transition, and description #24 (1–2). “Os quadratum, angle of proc. oticus and proc. orbitalis in lateral aspect (ordered): 0 large, 1 moderate, 2 small. CI ⫽ 0.67.” Comments: No difference in size of either orbital or otic process of quadrate between “Cu” and other cuckoos is apparent: proc. oticus and proc. orbitus in parasitic cuckoos Cuculus, Clamator and Urodynamis were as large as in nesting cuckoos of similar body size or larger (Ceuthmochares, Piaya melanogaster, Morococcyx, Centropus). Quadrate processes were larger in brood-parasitic Tapera than in the larger Morococcyx. There was a trend for larger cuckoos to have larger processes than smaller cuckoos, although this did not consistently distinguish groups, and quadrate processes varied in a gradual transitional series and not in a step function. #37 (0–1). “Os exoccipitale, shape (ordered): 0, flat or slightly raised; 1, produced into bulbous expansion raised above contour of os occipitale; 2, as in 1 but greatly expanded. CI ⫽ 1.00.” Comments:The bulbous projection from the skull base was not very large (2) in any cuckoos, and the birds in which it was relatively large included both parasitic cuckoos (Tapera, Dromococcyx, Clamator, Cuculus, Eudynamys, Urodynamis) and nesting cuckoos (Geococcyx, Zanclostomus, Dasylophus, Morococcyx, Coccyzus, “Saurothera” and both Coua). #80 (0–1). “Pelvis, relative widths of facies dorsalis (ordered): 0, narrowest point less than half the width between processes antitrochantericus; 1, approximately half; 2, more than half. CI ⫽ 0.67.” Comments: Seibel listed this as a synapomorphy for Cuculinae, less Eudynamys and Urodynamis. Hughes (1997) listed 0–1 as a synapomorphy for “Cu”, 1–2 as a synapomorphy for most “Cu” and 1–0 as a reversal from other “Cu” to Eudynamys and Urodynamis. I find all forms to be “0”, except “1” for brood-parasitic Tapera and Dromococcyx, and a nesting cuckoo Coccycua minuta.

64 The Cuckoos Table 4.5 contd. #87 (0–1). “Ilium, cross-section of margo medialis (Fig. 15) of crista iliaca dorsolateralis in axis dorsoventralis: 0, broad, forming smooth arc; 1, compressed into thin, blade-like shelf. CI ⫽ 1.00.” Comments:This character ⫽ PE10 in Seibel 1988, I find that all cuckoos are “0” except for the brood-parasitic Tapera and Clamator and the nesting Crotophaga which are “1”. #92 (0–2). (p) “Ilium, relative length of sync. ilioischiadica from crista caudalis fossa renalis to proc. marginis caudalis in axis rostrocaudalis, with reference to depth of recessus caudalis fossae: 0, intermediate between state 1 and 2; 1, sync. ilioischiadica long, recessus caudalis fossae present and deep; 2, sync. ilioischiadica short, recessus caudalis fossae shallow. CI ⫽ 0.67.” Comments: Seibel (1988) defined the character in terms of vertebral centrum width; Hughes defined it in terms of depth of the fossa renalis. Seibel found three transitions, none corresponding to those of Hughes. Seibel (1988: 64) described the character as problematic and nearly impossible to code within the Phaenicophaeinae, that is, the group related to Cu. In general aspect, birds with a longer pelvis have a longer ilioishiadic fusion H93. The character varies with the life style of cuckoos, a shorter ilioishiadic fusion in the arboreal cuculines; the structure is nearly identical in the parasitic Tapera and the nesting Morococcyx, while it is shorter in the small, terrestrial parasitic Dromococcyx. I find character 92 to have state 2 in the brood-parasitic Dromococcyx and in the nesting cuckoos Coccyzus, Coccycua minuta and Piaya melanogaster. #94 (1–2). “Sternum, cranial extent of apex carinae (ordered): 0, cranial tip of apex caudal to labrum externum of sulcus articularis coracoideus; 1, apex at same cranio-caudal level as sulcus; 2, apex cranial to sulcus. CI ⫽ 0.50.” Comments: I find no difference among cuckoos. #111 (0–1). “Humerus, extremitas proximalis humeri,“bulbous convexity” on facies cranialis just lateral to midpoint: 0, absent, 1, present. CI ⫽ 1.00.” Comments: I find no difference among cuckoos. #112 (0–2). “Humerus, position of fossa m. brachialis in cranial aspect: 0, mesial to center of corpus humeri; 1, far mesial; 2, at or lateral to meso-lateral center of corpus humeri. CI ⫽ 0.67.” Comments: Seibel found the transition from primitive cuckoos to occur in two places, one leading to the New World terrestrial cuckoos Geococcyx, Morococcyx and Neomorphus, the other to Coua. Seibel’s transitions and Hughes’s transitions occurred at different points on their phylogeny; Seibel’s transitions did not occur to define a clade “Cu”. Examination of UMMZ specimens supports Seibel’s transitions, except that the form of the fossa varies within Coua. I find all cuckoos to be similar except for Crotophaga, Geococcyx, and Coua caerulea, all with a lateral position. #124 (1–2). “Os carpi ulnare, shape of distal end: 0, blunt, expanded, or bulbous; 1, pointed, margo distalis 45⫿ to crus longum; 2, strongly pointed, approximately 30°. CI ⫽ 0.67.” Comments: Seibel defined three characters of the ulnare—his “cuneiform”—that appeared to be synapomorphies of all cuckoos; Hughes used these to define cuckoos and also recognized a new character, the shape of the distal end, a modification of Seibel’s character CU3. In parasitic cuckoos the ulnare has a more slender shape, but not in all; it is blunt in Eudynamys and especially in Tapera, whereas in nesting cuckoos it is narrow and pointed in Saurothera and somewhat narrow in Ceuthmochares and Dasylophus. The form seems to vary with behavior, blunt in terrestrial forms, and narrow and pointed in the more arboreal birds and the intercontinental migrants. In addition, Seibel (1988) reported the following characters to distinguish the clade Cuculinae (except Urodynamis, which fell outside the clade): S#57, PE5: (pectineal process, ⫽ tuberculum preacetabulare of Hughes 2000). I find no discrete variation in this process between the parasitic and non-parasitic cuckoos, although it is large in the terrestrial forms, as noted by Verheyen (1956). S#55 , UL1:0–2, ulna distal shape: Seibel had 6 states for the ulna; I find no distinct differences in ulna shape. S#52 , PE17, iliac crest shape and its anterior position relative to antitrochanter) Seibel (p. 63) found PE17 problematic and nearly impossible to code in Coccyzus, and PE17 individualistic and nearly impossible to code in Rhinortha. In Seibel (1988), PE17 character state B was coded “not as in A” with no independent description. I find the antitrochanter indistinct in cuckoos, whereas it is distinct in the goose and turkey (Howard 1929, Baumel 1993).

Morphology 65 the phylogenetic hypothesis of Hughes (2000), and the character is not unique to brood-parasitic cuckoos. The models of Seibel (1988) and Hughes (2000) of phylogeny among the cuckoo species and a single origin of brood parasitism depended on certain characters. Deconstruction of these characters reveals unquestioned assumptions and inconsistencies and leaves no reason to think that broodparasitic behavior evolved only once in the cuckoos. Also, several characters that supported their conclusion of a single origin of brood parasitism changed more than once in the skeletal phylogeny of cuckoos. The characters with a low consistency index are 24, 80, 95, and 124 (Hughes 2000). In addition, certain of these reported character states are not discrete, but rather they intergrade between extremes: these intergrading characters are 24, 37, 87, 93, 95, 110, 112, and 124 (Hughes 2000). Posso (2003) found the skulls of cuckoos to differ from the descriptions of Hughes, and his phylogenetic estimate of cuckoo relationships also differed from that of Hughes. There is no confidence in conclusions that are based on unrepeatable characters. The conclusion that brood parasitism evolved once rather than twice in the Old World cuckoos involves several of these same characters: the first skeletal estimate, based on postcranial characters (Seibel (1988). The same conclusion in Hughes (1997, 2000) involved 12 characters that changed at the node of separation of the malkohas from the Old World brood parasites. Several of these characters changed more than once in the skeletal phylogeny of cuckoos (following the numbers in Hughes (2000), these were characters 5, 33, 89, 113, 115) and had a low consistency index (CI). The other characters (4, 31, 80, 100, 108) were either not discrete, or had multiple states with questionable linear polarity, or had both these features (e.g. 66), or had a low CI and were not discrete (e.g., 24). In brief, this node is not well supported. Inasmuch as the authors did not test the relationships of cuckoos with other birds, the conclusion that Asian ground-cuckoos Carpococcyx are the basal cuckoos is a consequence of the assumption that character states in turacos are the prim-

itive character states in cuckoos. Because the assumption was not tested, and the relationship of cuckoos and turacos is not apparent in comprehensive systematic surveys of birds (Sibley and Ahlquist 1990, Livezey and Zusi 2001), the conclusions from the morphological estimates are open to question. In particular, the skeletal characters used in earlier studies do not support the conclusion of a single origin of brood parasitism in the cuckoos.

Other morphological variations among cuckoos Other morphological features that vary among cuckoos are the form of the syrinx, the arrangement of feather tracts, and the presence and absence of certain leg muscles (coded into alphabetic formulae that note the presence and or absence of these muscles). These characters are not closely associated with each other in cuckoos (Berger 1960). Anatomical studies of cuckoos should be continued, not only to add to the skeletal descriptions of Pycraft (1903), Seibel (1988) and Hughes (1997b, 2000), but also to confirm the characters and to add taxa to previous descriptions. Table 4.6 summarizes the earlier comparative morphological studies of cuckoos. Some characters in Berger (1960) were from earlier descriptions (e.g. Beddard 1885, 1898b), and some characters differed among these earlier descriptions. Characters in Table 4.6 are mainly from Berger’s Tables 4.2–4.4, with information on additional taxa from his text. The cuckoo syrinx has been described by Beddard (1895, 1898a,b, 1901, 1902) and Berger (1960). A bronchial syrinx has enlarged bronchi; a tracheobronchial syrinx has both enlarged bronchi and lower trachea, and the intrinsic muscles are attached to a bronchial semi-ring. Cuckoos with a tracheobronchial syrinx include Phaenicophaeus curvirostris (“erythrognathus”), Ceuthmochares, Scythrops, Eudynamys, Tapera, Clamator, Coccyzus, Saurothera, Piaya, Chrysococcyx, Cuculus, Hierococcyx Surniculus, Cacomantis and Pachycoccyx. Cuckoos with a bronchial syrinx Carpococcyx, Coua, Crotophaga and Guira; although Beddard (1885) indicated a tracheobronchial syrinx

66 The Cuckoos Table 4.6 Anatomical characters of cuckoos (after Nitzsch 1867, Beddard 1885, 1898a,b, 1899, Pycraft 1903, and Berger 1960). Character 1 2 3 4 5 6 7 8 9 Clamator 1 0 1 0b 1c 0 1 1 0 Cuculus 1 1 0a 0b 1c 0 0 1 0 Pachycoccyx 1 0d 0 “Surniculus” 1 0 0 0b 1c 0 1 1 0 a b Scythrops 1 1 0 0 1c 1 0 1 1 Eudynamys 1 1 0a 0b 1c 1 0 1 1 a b c Chrysococcyx 1 0 0 0 1 0 0 1 1 Ceuthmochares 1 0 0a 0b 1c 1 1 1 0 Phaenicophaeus 1 0 0a 0b 1c 1 1 1 1 Coccyzus 1 0 1 0b 0c 0 1 1 0 Saurothera 1 0 1 0b 0c 0 1 1 0 b c Piaya 1 0 1 0 0 0 1 1 0 Centropus 0 1 0 1 1 1 0 0 1 Carpococcyx 0 1 0 0 0 1 0d 0 1 Coua 0 0 0 0 0 1 1 0 1 Crotophaga 0 1 0 0 1 1 1 1 1 Guira 0 1 0 0 1 1 1 0 1 Tapera 1 0 0 0b 1 0 1 1 0 Dromococcyx 0 0 0 0 1 1 1d 0 1 Morococcyx 0 0 0 0 0 1 1 0 1 Geococcyx 0 0 0 0 0 1 1 0 1 1. syrinx: 0, bronchial; 1, tracheobronchial. 2. sternum: 0, double-notched; 1, single-notched. 3. cervical vertebrae: 0, 14; 1, 13 (aHughes found 13 only in Clamator and Coccyzus). 4. M. flexor hallicus longus inserts on hallux: 0, yes; 1, no (bBeddard 1898 reported it absent in Centropus, but Berger found it present, with a different attachment than in other cuckoos) (there was no specific description of the state in Tapera or other cuckoos, but it was inferred absent only in Centropus). 5. M. flexor metacarpi brevis: 0, absent; 1, present (csee Berger p. 77, also Berger’s Table 3). 6. M. piriformis ⫽ M. caudofemoralis, pars iliofemoralis “B”: 0, absent; 1, present. 7. M. iliacus “E”: 0, present; 1, absent (dPachycoccyx: absent in Berger’s Table 3, not mentioned p. 73; Dromococcyx: ? p. 73; (Scythrops: present in Berger’s Table 3; not listed p. 77). 8. dorsal apterium between cervical and interscapular tracts: 0, absent; 1, present. 9. ventral abdominal tracts: 0, single; 1, double.

in Guira.A few cuckoos have an intermediate syrinx, that is, a bronchial syrinx with some tracheal development: Phaenicophaeus curvirostris, Geococcyx, Morococcyx and Dromococcyx, and Centropus are reported to have either an intermediate or a bronchial syrinx (Beddard 1898b, 1902, Berger 1960). In fresh specimens of Carpococcyx renauldi (UMMZ 219043, 236029), a few details of feather tracts and leg muscles differ from those in

the descriptions of C. radiatus by Beddard (1901) and Berger (1960). I found that a dorsal median apterium was not present between the dorsal cervical tract and the spinal tract (there were scattered feathers as dense as in the interfemoral region of the spinal tract in Berger’s Fig. 3); there were two small apteria on the pelvic region of the spinal tract; and leg muscle “E” (M. iliacus) was present (Berger did not find it in a preserved specimen). In an alcoholic

Morphology 67 specimen of Rhinortha chlorophaea (USNM 223470) I found a single ventral feather tract on each side, like the single tract in Ceuthmochares aereus as described by Berger (1960), and in Zanclostomus javanicus (UMMZ 236439), Dasylophus superciliosus (UMMZ 228023, 2280244) and Piaya cayana (Beddard 1885), and not a partially double ventral feather tract as in Scythrops and Centropus as reported by Nitzsch (1867) or Phaenicophaeus curvirostris as reported by Beddard (1885). Studies of the bony palate, postcranial skeleton, leg muscles and the geographic distribution of the major groups of cuckoos show little agree-

ment between sets of data (Berger 1960, Seibel 1988, Hughes 1997d, 2000), and re-analysis revealed certain inconsistencies and questions in coding of the skeletal characters. Another estimate of cuckoo phylogeny (Hughes 1996a) used the occurrence of brood parasitism, egg color, behavioral and ecological traits. It is clear that there is no consensus about the relationships among the cuckoos from a comparison of these characters. It is unclear whether additional morphological and behavioral characters can provide a set of data that allows the same resolution as a good molecular data set.

5 A molecular genetic analysis of cuckoo phylogeny Perspective Historically, taxonomists organized the diversity of life into a system of names (or taxonomy) that reflected the general similarities and differences among different kinds of organisms.As information became available, the many species of cuckoos were recognized as belonging to a single avian family (or order) on the basis of the similarities in their behavior and morphology. Early taxonomists used a qualitative approach, in which species with similar appearance were grouped together. Systematists now agree that taxonomy should explicitly reflect the evolutionary relationships among organisms. Grouping species together in the family Cuculidae, for example, implies that they are descended from a single common ancestor that is more recent in time than the common ancestor of cuckoos and other more distantly related birds. Cuculidae also includes every species that is descended from their most recent common ancestor. In other words, each taxonomic group (e.g., order, family, subfamily, genus) is a natural or monophyletic group, and the hierarchical structure of these groups is consistent with an evolutionary tree or cladogram that depicts their historical relationships. Systematists use various kinds of information and analytical approaches to construct evolutionary trees. Overall similarity, quantified using either morphological or genetic data, can be used to establish relationships under the assumption that similarity between taxa is related to the time since common ancestry. This assumption, however, does not always hold, as evolutionary change may proceed at different rates in different lineages.

Although phenetic methods once were widely used and may be the only option for certain kinds of data (e.g., DNA-DNA “hybridization” or thermal dissociation), most systematists recognize that shared derived characters (novel morphological, behavioral, or genetic characteristics that are inherited from a common ancestor) provide the best evidence of common ancestry and in turn provide the basis for discovering the hierarchical relationships among organisms. Building a branching tree of life would be a trivial process if each character examined evolved only once and was retained in all the descendants of the ancestor in which the character first appeared, and if we could find at least one character to diagnose each group and subgroup in our tree. Similar features, however, may evolve independently in distantly related organisms and a derived character state may be lost in some descendants of the common ancestor in which it first appeared. In both situations, the organisms that share a particular character state may not represent a natural (monophyletic) group. Homoplasy, the independent origin or loss of a character state in unrelated organisms, obscures evolutionary relationships and is the essential problem that systematists confront in building trees. Systematists now use quantitative methods to select the tree or, equivalently, the hypothesis of evolutionary relationships that provides the best explanation for all the available character data. In a phylogenetic parsimony analysis, for example, the tree that requires the smallest number of evolutionary transformations to explain the distribution of character states in the terminal taxa is the preferred hypothesis of evolutionary relationships, because it

A molecular genetic analysis 69 minimizes the number of ad hoc hypotheses of homoplasy needed to account for the data. For example, if three different derived character states are shared by species A and B, and only one is shared by species A and C, we prefer the hypothesis that A is more closely related to B rather than to C.The former hypothesis requires five evolutionary transformations or steps (including one instance of homoplasy), whereas the latter requires seven steps (including three instances of homoplasy). Phylogenetic analyses can use any kind of discrete character data, including information from morphology, behavior, and molecular genetics. For a variety of reasons, systematists are relying to an ever greater degree on molecular genetic information and particularly DNA sequence data. DNA sequences offer a number of advantages, perhaps most importantly the availability of a very large number of potentially informative characters that are comparable across a broad diversity of taxa. As illustrated in the preceding section, morphological characters may be subject to a degree of ambiguity and interpretation associated with establishing the homology of characters across taxa and the scoring of characters, that may not fall neatly into discrete states. Perhaps more problematic is the potential for a suite of morphological characters associated with a particular lifestyle (e.g., arboreal versus ground-dwelling) to evolve in a correlated fashion in independent lineages, resulting in correlated homoplasy that can mislead a phylogenetic analysis. Molecular data is not immune from such problems, but it offers much less ambiguity in the initial scoring of characters (the nucleotide at any particular position in a DNA sequence must be either A, C, G, or T, the four nucleotides that comprise DNA) and they offer potentially greater character independence. Most differences in DNA sequences among closely related taxa are likely to reflect mutational changes that have little if any effect on the function of the genes involved and therefore should show little potential for correlated convergence associated with particular ecological or behavioral adaptations. Molecular data, however, is limited in the number of possible character states (four for DNA sequences), and is therefore characterized by levels of homoplasy that may be greater than that found in morphological data.As

a result, any one molecular character is likely to be inconsistent with the evolutionary history of a group of organisms, but hierarchical patterns reflecting common ancestry generally emerge when a large number of molecular characters are considered together in a single analysis. A few previous studies have used genetic data to address cuckoo relationships. These studies have used cytological, protein, and molecular sequence data and have used both phenetic and discrete character approaches to phylogenetic analysis. (1) As in other birds, chromosome numbers in cuckoos are large. The typical diploid number is 78–80, including 6 pairs of macro-chromosomes (Yamashina 1946, Ray-Chaudhuri 1967, 1973, de Lucca 1974, Waldrigues and Ferrari 1980, 1982, Waldrigues et al. 1983, Jensen 1980; other references are in Erritzoe 2000). The large number of avian chromosomes and the presence of barely visible micro-chromosomes that are difficult to count or identify, greatly complicates attempts to use chromosome numbers as evidence of systematic relationships. Nevertheless, the shapes of chromosomes may provide phylogenetic information.As an example, the karyotypic morphologies of the Common Cuckoo and the Diederik Cuckoo are more similar than either is to the Common Koel (RayChaudhuri 1973). While this result is consistent with the molecular analysis described in the next section, karyotypic studies of cuckoos are too few to support further inferences of their relationships. (2) Brush and Witt (1983) used electrophoresis of feather keratins as a genetic test of cuckoo relationships. Their similarity matrix suggested that (1) Old World and New World cuckoos are distinct groups, (2) Old World brood-parasitic cuckoos are most closely related to the malkohas, (3) this clade is the sister group of the coucals, (4) the New World Coccyzinae are most closely related to the anis, and (5) this clade is most closely related to the New World ground cuckoos (Payne 1997b). As in other studies, only a few species were compared and not all these were compared directly with each other. Nonetheless, these results are generally consistent with the analysis of cuckoo relationships presented below (except for a different placement of the coccyzine cuckoos).

70 The Cuckoos (3) In a broad survey of all birds based on thermal dissociation of species-pairs of DNA, Sibley and Ahlquist (1990) recognized six families of cuckoos: Cuculidae, including the Old World parasitic cuckoos, Old World malkohas, couas and Old World ground-cuckoos Carpococcyx; the coucals Centropodidae; the American cuckoos Coccyzidae; the anis Crotophagidae; the hoatzin Opisthocomidae; and the New World groundcuckoos Neomorphidae. Recognition of these six cuckoo families was based on an arbitrary criterion for delineating family level taxa, using genetic distances derived from thermal dissociation data.Their study was limited in not including any malkohas, couas, Old World ground-cuckoos, or any crested cuckoos Clamator, or New World brood parasitic cuckoos; and although their systematic scheme (Sibley and Ahlquist, Fig. 360) shows a coucal, no coucal appears in their thermal dissociation curves (Figs. 58, 76–84) or in their pooled-species analysis of cuckoos (Fig. 333). Their analysis (Fig. 333) shows Coccyzus (along with Piaya and Saurothera) and Cuculus (with other Old World brood parasites) as sister groups, and Geococcyx and Crotophaga (with Guira) as sister groups, results that are consistent with our analyses (see below). Other aspects of Sibley and Ahlquist’s phylogeny are inconsistent with our analyses and appear to have been based on limited data (inclusion of hoatzin in Cuculiformes and placement of coucals Centropus as the sister group to Cuculidae), or no data at all (inclusion of malkohas, couas, and Carpococcyx in Cuculidae). (4) In an analysis of partial sequences of the mitochondrial cytochrome b and ND2 genes from 15 cuckoo species, Aragón et al. (1999) found that brood-parasitic cuckoos occurred in three separate clades: (1) Old World brood-parasitic cuckoos (Cuculus, Cercococcyx, Cacomantis, Chrysococcyx), (2) malkohas (including an Old World nesting cuckoo Dasylophus superciliosus, New World Piaya cayana and the Old World brood parasites Clamator jacobinus and C. glandarius), and (3) New World anis and ground-cuckoos (Guira, Geococcyx, Neomorphus, along with the brood-parasites Tapera and Dromococcyx). Following Hughes (1996a), the authors suggested that nesting behavior in some cuckoos may have been derived from parasitic behavior,

although this conclusion does not follow from a parsimonious reconstruction of reproductive behavior on their phylogeny. Limited sampling of both molecular characters (851 base pairs per taxon) and cuckoo taxa, and the inclusion of a single outgroup species may all have contributed to somewhat ambiguous results in this study. (5) Johnson et al. (2000) analyzed mitochondrial DNA sequences for 26 cuckoos, the hoatzin, and two turacos, with a focus on the couas. Their results suggested that the Old World brood parasites are most closely related to the New World “Coccyzinae” (Coccyzus and Piaya) and that these two clades are in turn sister to the couas. Coucals Centropus were at the base of the tree within a clade also including the New World cuckoos Geococcyx, Guira, and Crotophaga.This study was also based on a limited set of molecular characters (951 base pairs per taxon) and did not include a number of key taxa, such as the Old World malkohas, Carpococcyx, and the brood-parasitic Tapera, Dromococcyx, and Clamator. In spite of some differences, the above studies show several results in common, and suggest the potential for molecular sequence data to provide a robust analysis of cuckoo systematics. All the above studies included a limited sample of taxa and therefore generate inferences primarily about relationships among the major groups of cuckoos. By necessity, a complete phylogenetic hypothesis that resolves relationships at the genus and species level requires the sampling of character information for all species. In addition, the two recent DNA sequencing studies on cuckoos have included a relatively small number of molecular characters per taxon, limiting the power of these analyses to resolve certain aspects of the phylogeny. In the present work, we have obtained a larger set of DNA sequence data for a complete sample of extant cuckoo species, to provide a robust and comprehensive analysis of the evolutionary relationships among cuckoos. As in many other avian systematics studies, our analysis is based on DNA sequences from the mitochondrial genome. The mitochondrial DNA (mtDNA) is a small genome, comprising about 17,000 base pairs of double-stranded DNA in a

A molecular genetic analysis 71 closed loop. It resides within the cell’s mitochondria and is independent of the chromosomal DNA in the nucleus.The mtDNA includes a small set of genes that is associated with cellular respiration and mitochondrial protein synthesis.The same complement of 13 protein-coding genes, 2 ribosomal RNA (rRNA) genes and 22 transfer RNA (tRNA) genes is shared by nearly all metazoa. A number of characteristics make mtDNA well suited for phylogenetic analyses of relatively closely related taxa. Because it is maternally inherited and effectively haploid, mtDNA has a smaller effective population size than nuclear genes, which results in a greater likelihood that the mitochondrial “gene tree” will correspond to the history of speciation events in the group under study (Avise 1994, Moore 1995, Klein and Takahata 2002). In addition, mtDNA exhibits a relatively rapid rate of evolutionary change, such that it provides ample information about relationships among closely related species. On the downside, this high rate of evolution results in a relatively high level of homoplasy in comparisons among more distantly related taxa. Particularly important for our study, mtDNA also has a high copy number per cell in avian tissues other than blood.This makes it feasible to amplify and sequence from samples with only small amounts of degraded DNA. Our goal of a comprehensive analysis of cuckoo relationships and the fact that many kinds of cuckoos have not been collected recently required that we obtain samples for many taxa from museum specimens that were collected up to 100 or more years ago. We find that mtDNA sequences can be obtained readily from most specimens less than 50 years old, and, by amplifying smaller segments of DNA, from specimens ranging up to 100 years or older (Payne and Sorenson, 2003).

Methods We extracted DNA from muscle tissue obtained from museum tissue collections or from the base of a feather plucked from live birds in the field or in captivity. For analyses of older museum skins, we extracted DNA from the base of a single feather or a small piece of skin cut from the foot. DNA extractions from museum skins were carried out in

a dedicated laboratory from which PCR products and extracts from fresh tissues were excluded. As in previous studies (Sorenson et al. 1999, Sorenson and Payne 2001), we amplified mitochondrial gene regions in overlapping fragments using the polymerase chain reaction (PCR) and sequenced the products using fluorescently labeled di-deoxy nucleotides and an automated DNA sequencer. For this study, we sequenced the mitochondrial genes for NADH dehydrogenase subunit 2 (ND2) and the small subunit ribosomal RNA (12S) along with portions of the transfer-RNA genes flanking both ends of both genes. Our choice of genes to sequence was intended to generate a large number of phylogenetically informative characters while also maximizing the potential independence of the data.We therefore chose two genes evolving under different functional constraints, such that possible systematic biases (Swofford et al. 1996) would be less likely to affect both genes in the same way. The mitochondrial cytochrome b gene has been sequenced more often than any other gene in avian molecular systematics studies to date (Sorenson et al. 1999), but the utility of this gene has been questioned (Meyer 1994, Weins and Hollingsworth 2000), and recent analyses suggest that other, more variable genes, provide greater phylogenetic signal (Yang 1998, Klicka et al. 2000).Among avian taxa for which complete mitochondrial genomes have been sequenced (Desjardins and Morais 1990, Harlid et al. 1997, Mindell et al. 1999), all protein-coding genes show similarly high levels of variation in third-codon positions, but vary greatly in amino acid conservation. Our choice of ND2 was based on the availability of robust avian-specific primers for this gene (Sorenson et al. 1999), and the expectation that it would provide greater phylogenetic information, particularly in the first and second positions, than other mitochondrial protein-coding genes (ND2 is the third most variable gene in amino acid sequence, after ATPase 8 and ND6, both of which are much shorter in length; Saccone et al. 1999). An important advantage of the 12S gene for this study is the presence of many conserved sequence blocks associated with the secondary structure of the rRNA molecule. This allowed us to design primers for both DNA strands spaced

72 The Cuckoos at 200–300 base intervals and to amplify the gene in small fragments when working with degraded DNA samples from older museum specimens. Obtaining ND2 sequences from older samples was more difficult because primers designed to amplify smaller fragments had to be tailored for different groups of cuckoos. Table 8 lists the cuckoo samples included in our genetic analysis of cuckoo phylogeny. Our objective was to include samples from all cuckoo species, and we obtained at least some sequence data for 202 individual cuckoos representing all but one of the 141 species recognized here. Sequences of both ND2 and 12S are complete for 143 samples, representing 119 species. As in the analysis of skeletal characters, we also include in our analysis several outgroup taxa (not in Table 8): mousebirds Coliiformes (2 species), parrots Psittaciformes (2 species), turacos Musophagiformes (2 species), the hoatzin Opisthocomiformes, doves Columbiformes (2 species), songbirds Passeriformes (3 species), a duck (Anseriformes), chicken (Galliformes), and 2 ratites (Struthioniformes). The entire data set comprises 399,155 base pairs of sequence information. For samples with complete data, the 12S and ND2 sequences (along with portions of the flanking tRNA genes), respectively, include 1020– 1042 and 1085–1098 base pairs of sequence data per taxon. In a preliminary alignment of these sequences, 737 positions of the 12S region and 1074 positions of the ND2 region (including the entire coding sequence) align unambiguously and without gaps in any taxa. Given that a significant portion of the 12S data set comprises sequence regions that are variable in length among taxa, and given that much of the informative variation in this gene occurs in these regions (Sorenson and Payne 2001), sequence alignment was a very important issue in our analysis.Typically, molecular systematists align sequences as the first step in a phylogenetic analysis. Where sequences vary in length among taxa, gap characters (“-”) are introduced as necessary so that nucleotides that are homologous across taxa are included in the same columns of the data matrix. For a set of closely related taxa such as the cuckoos, it is relatively easy to identify and align conserved portions

of the sequence, either manually or with the aid of a computerized alignment algorithm (e.g., Clustal W; Thompson et al. 1994). Inevitably, however, variable regions remain for which the alignment of sequences is ambiguous, and these regions are typically excluded from phylogenetic analyses. Exclusion of this data makes sense if the arrangement of nucleotides into columns is essentially arbitrary, such that there is little confidence that the nucleotides in a column share positional homology. Unfortunately, these regions of uncertain alignment may contain a great deal of potential information about phylogenetic relationships (Giribet and Wheeler 1999, Sorenson and Payne 2001). We think the best solution to this problem is the optimization alignment approach described by Wheeler (1996) and implemented in the computer program POY (Gladstein and Wheeler 1996). In an optimization alignment tree search, sequences are input into the analysis without a priori alignment. In effect, the alignment is reconsidered (optimized) for each of the different tree topologies considered by the tree search algorithm, to find the alignment/topology combination that provides the most parsimonious explanation of the data. This approach leads to far more parsimonious solutions than the usual two-step process of creating a static alignment and then searching for the best or shortest tree for that particular alignment. It allows the information present in variable “gap regions” to be used in a rigorous and unbiased way. Based on preliminary analyses, we excluded 45 individual samples with sequences that were identical or nearly identical to another individual of the same species, leaving a set of 171 individuals (157 cuckoos and 14 outgroup taxa) for our principal analyses. Our analysis of this data set in POY closely followed the approach used by Sorenson and Payne (2001). For each of two weighting schemes, we completed 50 replicate tree searches, each with a different random addition of sequences to the analysis. In the first set, all changes (transitions, transversions, insertions and deletions) were given equal weight. In the second analysis, transitions, (changes between the two purines, A and G; or between the two pyrimidines, C and T), were given a weight of 1, whereas transversions (changes between purines

Table 5.1 List of specimens used for DNA sequencing. The museum specimen number is provided for genetic samples that are associated with a voucher specimen. Note that one or more additional numbers identifying the collector, preparator, and/or tissue collection are often associated with these specimens. Tissue types included T: samples from frozen tissue collections (generally muscle); M: a feather and/or skin sample from museum skin specimens; and F: feather samples for which no voucher specimen is available (e.g., samples taken from live birds in captivity or birds that were captured and released in the field). Sequence data collected to date from the mitochondrial small subunit ribosomal RNA (12S) and NADH dehydrogenase subunit 2 (ND2) genes are indicated for each sample (complete sequence: √; partial sequence: 0.5, 0.75). Locality

Guira guira Guira guira Crotophaga major Crotophaga ani Crotophaga sulcirostris Tapera naevia Dromococcyx phasianellus Dromococcyx pavoninus Morococcyx erythropygus Geococcyx californianus Geococcyx velox Neomorphus geoffroyi squamiger Neomorphus geoffroyi salvini Neomorphus radiolosus Neomorphus rufipennis Neomorphus pucheranii Centropus milo Centropus ateralbus Centropus menbeki Centropus menbeki Centropus chalybeus Centropus unirufus Centropus chlororhynchos Centropus chlororhynchos Centropus celebensis celebensis Centropus celebensis celebensis

Bolivia, Santa Cruz Dept. captive Ecuador Puerto Rico, Cabo Rojo Nicaragua Bolivia, Santa Cruz Dept. Peru, Ucayali Dept. Peru, Ucayali Dept. Nicaragua Arizona, Pima Co. Nicaragua Brazil, Amazonas, Parintins, rt bank R. Amazon Panama, Darién Prov. Colombia, Cauca Dept. Guyana Peru, S. Rio Amazonas Solomon Islands, New Georgia Bismark Archipelago, New Ireland Indonesia,West Irian, Koesik Indonesia,West Irian, Lake Sentani Biak Philippines, Luzon Sri Lanka Sri Lanka Sulawesi Sulawesi

Museum Specimen No.

Tissue Type LSU 125706 (skeleton) T UMMZ 236027 T ANSP 83032 T LSU 150033 T UNLV 4308 T LSU 150544 T LSU 156873 (partial skeleton) T LSU 156024 T UWBM 69016 T UMMZ, feathers only F UWBM 68990 T AMNH 278613 M LSU 108160 T YPM 54524 M ANSP (skin in Guyana museum) T LSU 114631 T UWBM 63013 T ZMUC Noona 17.04.1962 M RMNH 4268 M YPM 74879 M ANSP 132995 M DMNH 68328 M MCZ 184954 M YPM 42613 M MCZ 270146 M RMNH 61684 M

Sequences 12S ND2 √ √1 √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ 0.5 √ √ √ √ 0.5 √ √ √

A molecular genetic analysis 73

Species

Species

Locality

Museum Specimen No.

Centropus rectunguis Centropus melanops Centropus steerii Centropus nigrorufus Centropus s. sinensis Centropus s. andamanensis Centropus toulou Centropus phasianinus nigricans Centropus phasianinus spilopterus Centropus phasianinus phasianinus Centropus bernsteini Centropus goliath Centropus goliath Centropus violaceus Centropus grillii Centropus viridis viridis Centropus bengalensis lignator Centropus bengalensis javanensis Centropus senegalensis “epomidis” Centropus senegalensis Centropus leucogaster leucogaster Centropus leucogaster efulensis Centropus anselli Centropus monachus Centropus cupreicaudus Centropus superciliosus burchellii Centropus superciliosus loandae Centropus superciliosus sokotrae Carpococcyx viridis

Malay Peninsula Philippines, Mindanao Philippines, Mindoro Java India, Haridwar Andaman Is Madagascar New Guinea, Morobe Prov. Kei Islands,Toeal Australia, QD, Shoalwater Bay Papua New Guinea, Maiwara Halmahera Halmahera Bismark Archipelago, New Ireland Malawi, Mchinji Dist. Philippines, Negros Taiwan Philippines, Negros Nigeria, Lagos Gambia Liberia Cameroon Angola, Cuanza Norte Prov. Angola, Malanje Prov. Zimbabwe South Africa, KwaZulu/Natal Namibia, Caprivi Strip Socotra Island Sumatra

AMNH 628149 CMNH 35734 YPM 34670 USNM 219367 UMMZ, feather only BMNH 87.12.2.1445 FMNH 384682 MVZ 149111 AMNH 628276 QM 29481 YPM 89146 MCZ 277981 USNM 571638 ZMUC Noona 17.04.1962 NMM 1988.1.34 CMNH 37080 MVZ 140273 CMNH 37079 BMNH 1951.7.8 UMMZ 235198 BMNH 1977.20.218 BMNH 1951.34.240 YPM 50356 YPM 50358 UMMZ, tissue, photo UWBM 62986 UMMZ 202732 USNM 518004 RMNH 04/EB

Tissue Type M T M M F M T M M M M M M M M T M T M T M M M M T T M M M

Sequences 12S ND2 √ 0.75 √ √ √ √ √ 0.5 √ √ 0.25 √ √ √ √ 0.75 √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ 0.75 √ √ √ √ √ √ √ √ 0.5 √ √ √ √ √ √ √ 0.75 0.15 √

74 The Cuckoos

Table 5.1 contd.

captive captive Sabah Madagascar Madagascar Madagascar Madagascar Madagascar Madagascar Madagascar Madagascar Madagascar Madagascar Thailand Sumatra Sumatra Central African Republic South Africa India, New Delhi Java, Jakarta market captive Malaysia, Sarawak Sri Lanka, Uva captive Sumatra India,Tamil Nadu India, Assam Thailand Philippines, Luzon Philippines, Luzon Philippines, Luzon Sulawesi Sulawesi

UMMZ 236029 UMMZ 219043 BPBM 10588 FMNH 352802 FMNH 384680 FMNH 356640 (skeleton) AMNH 628959 YPM 69417 FMNH 352799 FMNH 352797 FMNH 384681 FMNH 356638 (skeleton) MCZ 84298 OU 15103 MCZ 177724 YPM 42608 AMNH 10824 (alcoholic) DM 36943 MSU 4063 UMMZ, feather only UMMZ 236439 YPM 51093 FMNH 278258 UMMZ 236628 ANSP 139384 FMNH 229803 UMMZ 142974 OU 15102 CMNH 636502 UMMZ 233062 UMMZ 228024 RMNH 61661 MCZ 270138

T M M T T T M M T T T T M M M M T M M F T M M T M M M M T M M M M

√ —2 √ √ √ √ √ √ √ √ √ √ 0.25 √ √ √ √ √ √ √ √ √ √ √ 0.75 √ √ √ √ —2 √ √ 0.25

√ √ √ √ √ 0.75 √ √ √ √ √ √ 0.5 √ √ √ √ √ 0.75 0.5 √ 0.25 0.5 0.75 √ √ √ √

A molecular genetic analysis 75

Carpococcyx renauldi Carpococcyx renauldi Carpococcyx radiatus Coua caerulea Coua cristata Coua ruficeps olivaceiceps Coua verreauxi Coua coquereli Coua cursor Coua reynaudii Coua gigas Coua serriana Coua delalandei Rhinortha chlorophaea Rhinortha chlorophaea Rhinortha chlorophaea Ceuthmochares aereus Ceuthmochares australis Taccocua leschenaultii Zanclostomus javanicus Zanclostomus javanicus pallidus Phaenicophaeus sumatranus Phaenicophaeus pyrrhocephalus Phaenicophaeus curvirostris curvirostris Phaenicophaeus diardi Phaenicophaeus viridirostris Phaenicophaeus tristis “tristis” Phaenicophaeus tristis “longicaudatus” Dasylophus cumingi Dasylophus cumingi Dasylophus superciliosus Rhamphococcyx calyorhynchus Rhamphococcyx calyorhynchus

Species

Locality

Clamator coromandus Clamator glandarius Clamator jacobinus serratus Clamator jacobinus pica Clamator levaillantii Coccycua minuta Coccycua pumila Coccycua cinerea Coccycua cinerea Piaya melanogaster Piaya cayana Coccyzus melacoryphus Coccyzus minor Coccyzus ferrugineus Coccyzus euleri Coccyzus americanus Coccyzus erythropthalmus Coccyzus lansbergi Coccyzus pluvialis Coccyzus pluvialis Coccyzus rufigularis Coccyzus vieilloti Coccyzus merlini Coccyzus longirostris Coccyzus vetula Pachycoccyx audeberti Microdynamis parva grisescens Eudynamys scolopacea alberti Eudynamys scolopacea melanorhynca

Java, Jakarta market Gambia South Africa India, Haridwar Zimbabwe Bolivia, La Paz Dept. Colombia Bolivia, Beni Dept. Argentina Guyana Bolivia, Santa Cruz Dept. Ecuador Venezuela Cocos Island Ecuador Michigan Michigan Ecuador Jamaica Jamaica Dominican Republic Puerto Rico, San German Cuba, La Guira NP Dominican Republic Jamaica Central African Republic Indonesia,West Irian Solomon Islands Sulawesi

Museum Specimen No.

Tissue Type UMMZ, feather only F UMMZ, feather only F UNLV MBM 7489 T UMMZ, feather only F UMMZ, feather, photo F LSU 101256 (skeleton) T CU 35069 M LSU 123537 T USNM 614640 (wing, skeleton) T USNM 621716 (skeleton) T LSU (skin in Bolivia) T ANSP 187026 T MHNLS 10097 T CM 123814 M ANSP 185132 T UMMZ 225060 T UMMZ 236139 T ANSP 186069 T MVZ 149905 M UF 26233 M AMNH 164128 M LSU (skin in Puerto Rico) T ANSP 186545 T UMMZ, feather only F AMNH 155237 M AMNH 831777 T UMMZ 114159 M UWBM 58734 T MCZ 39733 M

Sequences 12S ND2 √ √ √ √ √ √ 2 — √ √ √ √ √ √ √ √ √ 2 — √ √ √ √ √ √ √ √ √ 0.75 0.75 √ √ √ √ √ √ √ √ √ √ √ √ 0.5 √ √ √ √ √ √ √ √ √ √ √ √ √ √ √

76 The Cuckoos

Table 5.1 contd.

India, Haridwar Australia, NSW Australia, QD New Zealand, Hamilton Australia, QD Thailand, Lampang Prov. Thailand,Tak Prov. Sumatra, Padang South Africa Uganda Central African Republic Kenya, Kericho Ghana Liberia Indonesia,West Irian Australia, NSW Australia, QD, Cape York Peninsula Australia, QD, Gowrie Ck/Herbert R junction Australia, QD, Cockatoo Ck Tanimbar New Caledonia Australia, NSW Indonesia,West Irian Papua New Guinea, Baiyer River Papua New Guinea,Wau Australia, NSW New Guinea Java, Jakarta market Java, Jakarta market Nepal Indonesia, Bacan Philippines, Mindanao Philippines, Mindanao

UMMZ, feather only CU 39830 QM 29927 AM 98.55 QM 29923 YPM 39592 YPM 68155 UWBM 67472 UWBM 53071 FMNH 364198 AMNH 24783 (skeleton) NMGL 1986.2.107 LSUMZ 168425 AMNH 827449 YPM 75000 UWBM 57616 ANWC 28619 QM 28434 QM 27712 RMNH P4503 UMMZ 221890 UWBM 57460 YPM 74870 YPM 89142 YPM 89143 UWBM 57432 MVZ 149109 UMMZ feather, photo UMMZ feather, photo FMNH 275817 YPM 74832 FMNH 357425 FMNH 357424

F M M M M M M T T T T M T T M T T M M M M T M M M T M F F M M T T

√ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √



√ √ √ √ √ √ √ √ √ √ 0.25 √ √ √ √ √ √

√ √ √ √

√ √

A molecular genetic analysis 77

Eudynamys scolopacea scolopacea Eudynamys scolopacea cyanocephala Eudynamys scolopacea cyanocephala Urodynamis taitensis Scythrops novaehollandiae Chrysococcyx maculatus Chrysococcyx xanthorhynchus Chrysococcyx xanthorhynchus Chrysococcyx caprius Chrysococcyx caprius3 Chrysococcyx klaas Chrysococcyx cupreus Chrysococcyx cupreus Chrysococcyx flavigularis Chrysococcyx megarhynchus Chrysococcyx basalis Chrysococcyx osculans Chrysococcyx minutillus poecilurus Chrysococcyx minutillus barnardi Chrysococcyx minutillus crassirostris Chrysococcyx lucidus layardi Chrysococcyx lucidus plagosus Chrysococcyx ruficollis Chrysococcyx meyeri Chrysococcyx meyeri Cacomantis pallidus Cacomantis leucolophus Cacomantis sonneratii Cacomantis merulinus lanceolatus Cacomantis passerinus Cacomantis variolosus “heinrichi” Cacomantis variolosus sepulcralis4 Cacomantis variolosus sepulcralis

Species

Locality

Museum Specimen No.

Cacomantis variolosus sepulcralis Cacomantis variolosus variolosus Cacomantis castaneiventris Cacomantis flabelliformis Cercococcyx mechowi Cercococcyx mechowi Cercococcyx olivinus Cercococcyx montanus Surniculus dicruroides Surniculus dicruroides Surniculus velutinus velutinus Surniculus velutinus velutinus Surniculus musschenbroeki Surniculus lugubris Surniculus lugubris Hierococcyx vagans Hierococcyx varius Hierococcyx varius Hierococcyx sparverioides Hierococcyx hyperythrus Hierococcyx pectoralis Hierococcyx pectoralis Hierococcyx nisicolor Hierococcyx nisicolor Hierococcyx nisicolor Hierococcyx fugax Cuculus clamosus clamosus Cuculus clamosus gabonensis5 Cuculus clamosus gabonensis6

Borneo, Sabah Australia, QD, Brisbane Australia, QD, Claudie River 3rd Crossing Australia, QD, Maleny Central African Republic Uganda Angola, Cuanza Norte Prov. Uganda Nepal Bhutan Philippines, Mindanao Philippines, Mindanao Halmahera Sabah Sabah Thailand India, Assam Oman Nepal Japan, Nagano Pref. Philippines, Mindanao Philippines, Mindanao India, Darjeeling Dist. Singapore Singapore Singapore South Africa Nigeria, Ibadan Nigeria, Ibadan

Sabah Parks 17065 QM 26747 QM 27099 QM 30248 AMNH 831292 UMMZ 221537 YPM 50330 FMNH 355263 FMNH 221725 USNM 519749 FMNH 286173 YPM 61680 USNM 571628 WFVZ 37595 WFVZ 37594 USNM 534592 UMMZ 142758 BMNH 1989.5.1 AMNH 831293 UMMZ, skin lost in Japan FMNH 357420 CMNH 35841 MCZ 297472 UWBM 67458 UWBM 67545 UWBM 67529 UWBM 52912 CU 29762 CU 29759

Tissue Type T M M M T M M T M M M M M M M M M M T T T T M T T T T M M

Sequences 12S ND2 —2 √ √ √ √ √ √ √ √ √ —2 0.5 √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ 0.75 √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ 0.5 0.5

78 The Cuckoos

Table 5.1 contd.

Cameroon South Africa Zambia Japan Vietnam,Tonkin Sulawesi Singapore Madagascar South Africa,Transvaal Botwswana Japan Philippines, Luzon Philippines, Panay Philippines, Panay Java, Jakarta market Nepal Philippines, Luzon China, Xinjiang UK Japan Russia Russia

FMNH 188943 UWBM 52799 UMMZ, feather only UMMZ 234916 AMNH 833647 AMNH 298824 UWBM 64917 FMNH 384675 UMMZ 216724 USNM 527286 UMMZ 234743 CMNH 36498 CMNH 36760 NMP (tissue: CMNH 4334) UMMZ feather only AMNH 831292 FMNH 6025 (voucher missing) CAS 98001 UMMZ, ex Univ. East Anglia UMMZ tissue only UWBM 46350 UWBM 49796

M T F T T M T T M M T T T T F T T M T T T T

√ √ —2 √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √

0.5 √ √ √ √ √ √ √ √ √ √ √ √

captive captive Paraguay captive Gambia Gambia Florida Japan Cameroon Gambia

UMMZ UMMZ UMMZ UMMZ UMMZ UMMZ UMMZ UMMZ UMMZ UMMZ

T T T T F F T T T F

√ √ √ √ √ √ √ √ √ √

√ √ √ √ √ √ √ √ √ √

234219 233646 227492 232536 (A573) (A794) 225650 234875 232516 (A525)

√ √ √ √ √ √ √

A molecular genetic analysis 79

Cuculus clamosus gabonensis Cuculus solitarius Cuculus solitarius Cuculus poliocephalus Cuculus poliocephalus Cuculus crassirostris Cuculus micropterus Cuculus rochii Cuculus gularis Cuculus gularis Cuculus optatus Cuculus optatus Cuculus optatus Cuculus optatus Cuculus saturatus or c. optatus 7 Cuculus saturatus Cuculus saturatus 8 Cuculus canorus subtelephonus Cuculus canorus canorus Cuculus canorus canorus Cuculus canorus canorus Cuculus canorus canorus Outgroup Taxa Colius striatus Urocolius macrourus Nandayus nenday Neophema elegans Musophaga violacea Crinifer piscator Columba leucocephala Treron sieboldii Vidua chalybeata Lanius senator

Species

Locality

Museum Specimen No.

Sequences Tissue Type 12S ND2 Smithornis sharpei Nigeria UMMZ 233887 T √ √ Sayornis phoebe Louisiana UMMZ 226790 T √ √ Opisthocomus hoazin Bolivia LSUMZ 131920 T √ √ Opisthocomus hoazin Peru LSUMZ 156874 (skeleton) T √ √ 1 The ND2 sequence for this sample may be a nuclear copy but it is closely related to the sequence from the other Guira sample. 2 The 12S gene was not sequenced for these samples because their ND2 sequences were identical or nearly so to another sample from the same species. 3 This specimen (tissue FMNH 2051) was misidentified as C. klaas in Johnson et al. (2000). 4 This specimen (tissue FMNH 6635) was misidentified as C. merulinus in Johnson et al. (2000). 5 Breast black, as in four of five specimens from this locality. 6 Breast rufous, as in one of five specimens from this locality, which is intermediate between the forest zone and Guinea woodlands. 7 The tail feather of this specimen was identified to one of those species. 8 This specimen was identified on the basis of its DNA sequence because we were unable to locate the voucher specimen.The bird was identified as H. (“C.”) fugax by the collector, and as H. (“C.”) vagans in Johnson et al. (2000). Its DNA sequence is similar to other C. saturatus and C. canorus. Both C. canorus and H. vagans can be excluded, however, because neither occurs in the Philippines (Dickinson et al. 1991) and the DNA sequence suggests that it is neither H. vagans or H. pectoralis (formerly considered a form of H. fugax). Identification to C. saturatus is based on a wing measurement of 181 mm.

80 The Cuckoos

Table 5.1 contd.

A molecular genetic analysis 81 and pyrimidines) and “indels” were given a weight of 2. On average, transitions occur much more frequently than other changes (Wakeley 1996) and therefore may be less reliable as evidence of historical relationships. This is not to suggest that a weighted analysis will necessarily produce a tree that is a better estimate of phylogeny. Rather, our interest in weighting the data was to explore the stability of the results with regard to a change in assumptions. In general, results for cuckoos from the two different weighting schemes were very similar. Because of the complexity of the alignment problem, POY uses heuristic methods to estimate tree length (i.e., the number of evolutionary changes or steps required to explain the data), and therefore the tree length score assigned to any tree, particularly in a large analysis such as this, includes a small but unknown over-estimation. This makes it difficult to conclude that one particular tree (or trees) is the most parsimonious. We therefore considered not only the tree(s) with the lowest estimated score, but also trees within five steps of the best score. In the unweighted analysis, 11 trees with 16119 to 16124 steps were found in 9 of the 50 tree search replicates. In the weighted analysis, 10 trees with 20419 to 20424 steps were found in 7 of the 50 replicates.We used these 21 “best” trees to construct a majority rule consensus tree in which only those groups found in a majority of the 21 trees are shown.We use this consensus tree from the 171-taxon analysis as our primary hypothesis of cuckoo relationships in the discussion and figures that follow, but also point out alternative relationships found in one or more of these 21 best trees.We also conducted 100 replicate tree searches under each weighting scheme for five subgroups within Cuculidae, comprising respectively, (1) Crotophaginae plus Neomorphinae, (2) Centropodinae, (3) Couinae, (4) Phaenicophaeini, and (5) Cuculini.Analyses of these smaller sets of taxa allowed us to more thoroughly explore tree space within each subgroup with less ambiguity in tree scores. In general, however, results of these analyses were highly congruent with the overall analysis. Calculation of bootstrap values (Felsenstein 1985), which measure the relative support for different branches in the tree, was based on a standard static alignment and excluded 138 characters in

regions of particularly ambiguous alignment. In bootstrapping, a large number of replicate data sets are generated by randomly sampling characters from the original data set and then determining the proportion of replicate data sets that support each node in the tree.

Relationships, evolution and biogeographic histories The molecular data provides strong support for the monophyly of cuckoos and for the other avian orders in our analysis, respectively, but it provides only limited support for higher-level (inter-ordinal) relationships. None of our analyses indicated the inclusion of hoatzin within Cuculiformes, as suggested by Sibley and Ahlquist (1990), or a sisterrelationship between cuckoos and hoatzin.We also found no evidence for a close relationship between turacos and cuckoos or between turacos and hoatzin, relationships that have been previously suggested (Hughes and Baker 1999). Cuckoos were monophyletic in all our analyses (Tree 1), a result consistent with a historical view of cuckoo systematics, insofar as one or more species in all the major groups of cuckoos were originally described as a Cuculus: Linnaeus in his 1766 edition of Systema Naturae listed 22 species in his genus Cuculus. We have Cuculus canorus as the “typical” cuckoo from an historical and Eurocentric point of view, and many Old World brood parasites and nesting malkohas such as Raffles’s Malkoha that were originally described as Cuculus. In addition, Blue Coua and Crested Coua were described as Cuculus caeruleus and C. cristatus by Linnaeus, Javan Coucal as Cuculus nigrorufus by Cuvier, Pheasant Coucal as Cuculus phasianinus by Latham, Madagascar Coucal as Cuculus Toulou by P. L. S. Müller, Lesser Coucal as Cuculus bengalensis by Gmelin, Philippine Coucal as Cuculus viridis by Scopoli, Senegal Coucal Cuculus senegalensis by Linnaeus, Guira Cuckoo as Cuculus Guira by Gmelin, Lesser Roadrunner as Cuculus velox by A. Wagler, and American Striped Cuckoo as Cuculus naevius by Linnaeus. Within the cuckoo, species molecular analysis provides a remarkably good resolution of evolutionary

82 The Cuckoos

88 100

99

93 89 71 100 59 90

99

87 97

99 63 90

89

8 52

87

80 90 52

70

20 100 54 90

64 61 61 100

50 92 69

Crotophaginae

Neomorphinae Centropodinae Couinae

Phaenicophaeni

100

Guira Crotophaga Tapera Dromococcyx Morococcyx Geococcyx Neomorphus Centropus Carpococcyx Coua Rhinortha Ceuthmochares Taccocua Zanclostomus Phaenicophaeus Dasylophus Rhamphococcyx Clamator Coccycua Piaya Coccyzus Pachycoccyx Microdynamis Eudynamys Urodynamis Scythrops Chrysococcyx Cacomantis Surniculus Cercococcyx Hierococcyx Cuculus

Cuculinae

Cuculini

100 96

Figure 5.1. Genus-level phylogeny of cuckoos (Cuculidae) based on mitochondrial DNA sequences. Obligate broodparasites are underlined.The tree shown is a majority rule consensus of the 21 most parsimonious and nearly most parsimonious trees found in optimization alignment analyses using two different weighting schemes (see text).The analysis included 157 cuckoo sequences, plus 14 outgroup taxa (not shown). Genus-level relationships were consistent among trees, except for six branches (shown as dashed lines) that were found in most, but not all trees.The proportion of the 21 “best” trees including each of these clades is indicated below each branch (underline, italics). Alternative arrangements of cuckoo subfamilies and genera within Cuculinae were found in a small proportion of the best trees and these are described in the text. Bootstrap percentages based on 500 replicate data sets are shown above each node. Monophyly of all genera recognized here was strongly supported.

relationships and strong support for most groups or clades in the tree. A phylogenetic hypothesis for cuckoo genera is shown in Figure 5.1.The basal division in this tree separates a clade of cuckoos found only in the New World, from a more species-rich clade that includes all the Old World cuckoos plus a few New World genera. We recognize five subfamilies as representing the main phylogenetic lineages among cuckoos. These are Crotophaginae and Neomorphinae in the New World, Centropodinae and Couinae in the Old World, and Cuculinae which has its origin and greatest diversity in the Old World, but includes a clade of New World cuckoos represented by Coccycua, Piaya, and Coccyzus (Figure 5.1). Some of the trees found in our optimization

alignment analyses included alternative arrangements of the five cuckoo subfamilies. In 62% of the 21 “best” trees, subfamily relationships were as shown in Figure 5.1. In 29% of the best trees, Couinae was the sister group to Cuculinae (this result was obtained only in the weighted analysis); and in 9% of trees, the Centropodinae/Couinae clade was the sister group of all other cuckoos (this result was found only in the equal weights analysis). Because the topology shown in Figure 5.1 was found most often and under both weighting schemes, we used it as our working hypothesis of cuckoo relationships. Sequence data from additional mitochondrial genes or nuclear genes, and more thorough analyses, are needed to further test this hypothesis and provide

A molecular genetic analysis 83 better resolution of some aspects of cuckoo phylogeny. The linear systematic treatment of cuckoo genera and species in this book is based on Tree 1 and on the species-level phylogenies of cuckoo subgroups. By convention, the descendant branch in each bifurcation with fewer taxa is considered “basal”, and these taxa are listed first in the linear sequence. Monophyly was used as a minimal criterion for recognition of taxa at each systematic level above the species. In principle, each node or branching point in the tree defines two clades which can be named in a hierarchical system, but only subfamilies and genera are consistently named in this review. Each genus recognized was monophyletic in all of our phylogenetic analyses, and this criterion emphasizes the common ancestry of the species within each genus, rather than the morphological distinctiveness of a set of similar species. In general, the genera we recognize are consistent with morphological and behavioral traits, with the relative degree of genetic divergence between lineages, and with traditional classifications of the cuckoos. In the following trees, we consider the specieslevel relationships among cuckoos and describe in greater detail the results of our phylogenetic analyses. The trees are drawn as phylograms, in which branch lengths are proportional to the minimum number of changes in DNA sequence that must have occurred in each lineage. The total branch length separating any two terminal taxa provides a rough measure of the genetic distance between them and in turn an indication of the time since they shared a common ancestor, although reconstruction of branch lengths using parsimony may substantially underestimate the number of mutation events that occurred in each lineage. In the trees for each cuckoo subfamily or tribe, we show the majority rule consensus tree from the overall analysis, but describe any alternative arrangements of taxa that were found either in the 21 best trees from the overall analysis or in the tree(s) with the best score in our analyses of cuckoo subgroups. Owing in part to the benefits of dense taxon sampling (Zwickl and Hillis 2002) and perhaps also to the spacing of cuckoo diversification events in history, the mitochondrial data provide a

remarkably good resolution of almost all aspects of cuckoo phylogeny above the species level, and strong support for most clades in the phylogeny. In addition, results were generally stable in relation to changing the weights applied to different kinds of character change (i.e., transitions versus transversions). Some of the remaining uncertainties in the cuckoo tree involve taxa with incomplete sequence data,whereas others will require data from additional genes to resolve. Genetic material for taxa with incomplete data came from older museum specimens, from which DNA sequences were more difficult to obtain. The basal cuckoo lineage includes two groups that occur only in the New World (Figure 5.2). Crotophaginae includes the cooperatively breeding Guira Cuckoo Guira guira and the anis Crotophaga in which birds in social groups lay in a common nest from which females remove the eggs and sometimes the nestlings of their group mates. Guira is the basal taxon in this set and Crotophaga ani and C. sulcirostris are sister species, as they are in a phylogenetic analysis of cranial skeletons of the Crotophaginae (Posso and Donatelli 2001). The other lineage, Neomorphinae, includes the roadrunners Geococcyx and ground-cuckoos Morococcyx and Neomorphus, which are monogamous and rear their own young. Phylogenetic relationships among the large ground-cuckoos Neomorphus in the molecular analysis are much the same as suggested by Haffer (1977) on the basis of geography and plumage pattern. Neomorphinae also includes the brood-parasitic cuckoos Tapera and Dromococcyx. The placement of these parasitic birds within this clade is strongly supported by the genetic data and indicates independent origins of obligate brood parasitism in New World and Old World cuckoos. In coucals Centropodinae, the molecular results indicate several lineages (Figure 5.3).A basal lineage includes Centropus menbeki and C. chalybeus in New Guinea and Biak, C. ateralbus in the Bismarck Archipelago, and C. milo in the Solomon Islands. Close relationships among these species are consistent with the known geological history of the New Guinea – Melanesia region. Although the details of this history are particularly complex (Pigram and Davies 1987), the Melanesian arc with the

84 The Cuckoos 100 100 99 100 86 100

50 changes

Guira guira (2) Crotophaga major Crotophaga ani 100 Crotophaga sulcirostris Tapera naevia Dromococcyx phasianellus 100 Dromococcyx pavoninus Morococcyx erythropygus Geococcyx californianus 100 Geococcyx velox 99 Neomorphus rufipennis 87 Neomorphus pucheranii 100 Neomorphus radiolosus Neomorphus g. geoffroyi 97 98 Neomorphus g. squamiger

Figure 5.2. Phylogeny of the New World cuckoos (Crotophaginae and Neomorphinae).This single “best” tree was found in all optimization alignment analyses. In this, and in the following figures, the trees are drawn as phylograms in which branch lengths are proportional to the number of character state changes on each branch, as inferred by parsimony reconstruction using ACCTRAN optimization in PAUP* (Swofford 2002). Bootstrap values are shown above each node.Two samples of G. guira had nearly identical sequences of the 12S gene, but somewhat divergent sequences for the ND2 gene. We suspect that a nuclear copy of the ND2 gene was amplified and sequenced for one of the two samples.

Bismarcks and Solomon Islands came into contact with New Guinea around 10 to 5 million years ago (Lee and Lawver 1995, Michaux 1998, Polhemus and Polhemus 1998), suggesting the maximum age for a common ancestor of Melanesian coucals, and the period when coucals could move from New Guinea into this region. Coucals are absent on Buka and Bougainville but occur on each side of this area, with two species in the Bismarck Archipelago, and one further south in the Solomons. This distribution gap may reflect an ancient extinction or a more recent historical extinction. In the Pleistocene when sea levels were lower, the land mass of Bougainville and most of the Solomons region was larger, presumably allowing coucals to move from the Bismarck Archipelago to the Solomons by way of Bougainville. Bougainville and most island groups in the Solomons were separated from each other by rising sea levels only over the past 10,000 years. Melanesian people have maintained a fully developed and widespread cultural complex in the region for the past 3,500 years (Wickler and Spriggs 1988, Spriggs 1997, Merriwether et al. 1999), and for at least 29,000 years people have occupied the region where cultural remains have been found at the northern end of the

Solomons on Buka Island just north of Bougainville. Indigenous mammals have disappeared on Buka over the past few millenia, after the human introduction of marsupial possums (Phalanger orientalis), rats (Rattus elegans, R. praetor), pigs and dogs (Flannery and Wickler 1990, Flannery and Roberts 1999), all which may have been predators of ground-cuckoos. Buff-headed Coucals Centropus milo occur on other Solomon Islands that never were connected to the large island “Greater Bougainville” (“Greater Bukida”) during low sea level times of the Pleistocene (Diamond et al. 1976, Spriggs 1997, Mayr and Diamond 2001). Circumstantial evidence therefore points to the absence of large coucals as a result of colonization, and growth of human population and associated animals, and yet the two known sites with avian fossils on St. Matthais and New Ireland have not revealed any species that are known only from fossils and do not still live in the area (Steadman and Kirch 1998, Steadman et al. 1999). The extinction of birds on the islands of northern Melanesia has not been as extensive as on smaller islands that are more remote from continental sources, islands such as Hawaii, Easter Island and Henderson Island (Steadman and Olson 1985, Steadman 1995, Mayr and Diamond 2001).

A molecular genetic analysis 85 Centropus milo C. ateralbus 99 C. menbeki (2) C. chalybeus C. unirufus C. chlororhynchos (2) 100 C. melanops 95 33 91 C. steerii C. rectunguis C. c. celebensis (2) 19 C. anselli 100 C. l. leucogaster C. l. efulensis 95 82 C. senegalensis C. senegalensis “epomidis” C. monachus 71 C. cupreicaudus 72 C. superciliosus sokotrae 79 99 C. superciliosus burchellii C. superciliosus loandae C. nigrorufus 100 C. s. sinensis C. s. andamanensis 40 C. toulou 87 C. goliath (2) C. grillii 100 100 C. viridis C. bengalensis lignator 98 C. bengalensis javanensis 32 C. violaceus C. bernsteini 67 C. p. nigricans 95 C. p. phasianinus 50 changes C. p. spilopterus 100

97

100

Figure 5.3. Phylogeny of Centropodinae (Centropus).The tree shown is the majority rule consensus from the overall analysis. Relationships within this genus were stable among analyses, with two exceptions. First, the positions of C. senegalensis and the anselli/leucogaster clade were reversed in weighted analyses. Second, C. goliath was sister to either C. toulou or to the violaceus/bernsteini/phasianinus clade. Bootstrap values are shown above each node.Two samples of C. menbeki, C. chlororhynchos, C. celebensis, and C. goliath, respectively, had nearly identical sequences.

A second coucal lineage consists of species that occur from Africa to southern Asia, the Philippines, the Greater Sundas, New Guinea and Australia. Within this group, the Philippine Centropus unirufus and the Sri Lanka C. chlororhynchos are basal, followed by a clade comprising Short-toed Coucal C. rectunguis from southeast Asia, and two additional shorttoed species from the Philippines, C. melanops and C. steerii (Figure 5.3). Next is C. celebensis from Sulawesi. These coucals have relatively short hallux claws, largely independent of their body size or habitat. Another clade comprises the African coucals Centropus leucogaster, C. anselli, C. senegalensis, C. monachus, C. cupreicaudus and C. superciliosus.

Sister relationships between the large forest coucals C. leucogaster and C. anselli and between C. cupreicaudus and C. superciliosus are strongly supported. The latter result differs from earlier views based on plumage, that regarded C. cupreicaudus, and C. monachus as a single superspecies (Snow 1978), but is consistent with a more recent proposal by Irwin (1985). C. senegalensis (including the plumage phase “epomidis”) was sister to a clade comprising C. monachus, C. cupreicaudus and C. superciliosus in the equal weights analysis, but was the basal lineage in this African clade in the weighted analysis. The broadly distributed Greater Coucal Centropus sinensis and the Javan Coucal C. nigrorufus comprise

86 The Cuckoos another well-supported clade. Finally, we find a geographically and morphologically diverse clade including the Madagascar Coucal C. toulou; two island species with large body size, C. goliath and C. violaceus; a clade comprising the African Black Coucal C. grillii, the Philippine C. viridis, and the widely distributed C. bengalensis, a small bird but with the longest hallux claw of any coucal; and a well-supported group comprising Pheasant Coucal C. phasianinus and Lesser Black Coucal C. bernsteini of New Guinea. Alternative placements of C. goliath were sister to C. toulou or sister to a clade comprising C. violaceus, C. bernsteini, and C. phasianinus, the latter seemingly more consistent with morphology and biogeography. Historically the coucals within a geographic region have been grouped by systematists in terms of plumage and size. Salvadori (1881) recognized three genera in Sulawesi and the Moluccas to New Guinea: (1) Nesocentor (Cabanis and Heine, 1863) (the large species goliath, milo, violaceus, menbeki, chalybeus and ateralbus), (2) Polophilus (Leach, 1814) (including the forms spilopterus, nigricans and bernsteini) and (3) Centrococcyx ( javanensis ⫽ bengalensis, rectunguis), the black-and-brown coucals.Then after a period when several genera were recognized, the coucals were again collapsed into a single genus Centropus (Hartert 1891, Shelley 1891). Mason et al. (1984) and Mason (1997b) retained a single genus but recognized three groups in Australasia based on juvenile plumage and habitat. These three groups corresponded in part with Salvadori’s scheme, although they were not based on phylogenetic analysis and they did not include all coucal species of the Australasian region. The genetic results lead to very different conclusions about the relationships among coucals and indicate that features of body size and plumage color have evolved repeatedly. For example, the large black coucals Centropus goliath and C. phasianinus spilopterus are in the same lineage as most of the smaller black-andbrown coucals, some of which have a seasonal change of plumage (C. toulou of Madagascar, C. grillii of Africa, C. bengalensis of Asia) while others do not (C. viridis of Asia, C. goliath). This result is consistent with the similarity in size and dull black plumage between C. goliath and Pheasant Coucal C. p. spi-

lopterus of the Kai Islands; these large coucals are also similar in the pale patches on their wing coverts. Large body size in C. goliath and C. p. spilopterus, however, reflects a general pattern of body size evolution in island birds, much as in the large subspecies of Lesser Coucal on islands (C. bengalensis sarasinorum on the Lesser Sunda Islands, C. b. medius on the Moluccas), of C. p. mui on Timor, and of Coccyzus lizard-cuckoos in the West Indies. Large body size does not necessarily indicate phylogenetic relationship, insofar as the large C. violaceus and C. milo of the Bismarck Archipelago and Solomon Islands are not closely related, nor are C. goliath of the northern Moluccas and C. menbeki of New Guinea. Couinae includes as sister groups the Madagascar couas Coua and the southeast Asian ground-cuckoos Carpococcyx (Figure 5.4).The two genera are similar in plumage, nests, and the naked skin and brightly patterned mouths of nestlings. Our results for species-level relationships among couas are consistent with those from a recent molecular study ( Johnson et al. 2000); our analysis includes four additional species. The two coua species living in open country, Verreaux’s Coua C. verreauxi and Crested Coua C. cristata, are closely related to each other.The extinct Snail-eating Coua Coua delalandei is most closely related to the Giant Coua C. gigas or Red-breasted Coua C. serriana, which it resembles in size and in dark rufous plumage. Placement of C. delalandei was uncertain because to date we have obtained only 334 base pairs of the 12S sequence from the ~140 year old specimen that we sampled. Within the ground-cuckoos Carpococcyx, relationships are not well-resolved because we obtained only 174 base pairs of the 12S sequence from an old specimen of Sumatran Ground-cuckoo C. viridis. The data nonetheless suggests that C. viridis is distinct as a species in that its 12S sequence differs from that of C. radiatus and C. renauldi by 5% and 9%, respectively. On the basis of genetic distance, C. viridis and the Bornean Ground-cuckoo C. radiatus may be more closely related to each other than either is to the Coral-billed Ground-cuckoo C. renauldi. The common origin of Madagascar couas and Asian ground-cuckoos may date from a time when these land areas were in contact or were closer together than at present. To understand the

A molecular genetic analysis 87 Carpococcyx viridis 100

Carpococcyx radiatus 65

Carpococcyx renauldi (2)

100

Coua cristata 100

Coua caerulea

96

Coua verreauxi

100

Coua ruficeps olivaceiceps 98

Coua reynaudii 99

100

Coua coquereli Coua cursor

47

Coua gigas 98 54 50 changes

Coua delalandei Coua serriana

Figure 5.4. Phylogeny of Couinae (Carpococcyx and Coua), with bootstrap values shown above each node. Relationships within this group are strongly supported with the exception of Coua reynaudii which is sister to the C. gigas/delalandei/ serriana clade in one of the 21 “best” trees. In addition, placements of Carpococcyx viridis and Coua delalandei are not well-supported because limited data was available for these taxa (149 and 334 base pairs of sequence data, respectively).The short terminal branches for these two species, and the seemingly close relationship between C. delalandei and C. serriana are artifacts of missing data. C. delalandei is only slightly more similar to C. Serriana than to C. gigas. Two samples of C. renauldi had nearly identical sequences.

distribution of land birds it is necessary to know the history of land and sea, and the distance between the land areas above sea level (Hall 1998).The geological and climatic history of the Indian Ocean region and the distribution of other terrestrial vertebrates suggest that historically more land was exposed and the sea itself was less of a barrier to dispersal and colonization of land animals than it is now. Madagascar and the drifting Indian subcontinent separated c. 130 mya (Krause et al. 1999), and the Indian subcontinent accreted with Eurasia c. 65–55.5 mya with a suture complete by 49 mya (Beck et al. 1995, Gower et al. 2002). The granitic Seychelles in the Indian Ocean between India, Madagascar and Africa are of continental origin and separated from Africa around 75 mya; the Indian Ocean has been in place between India and Madagascar for at least 50 million years and is very deep (Hall 1998). The Seychelles are emergents of a much larger land mass, the Seychelles Microcontinent, a now largely submerged, continental fragment in the Indian Ocean, about 55,000 km2

in area. This land would have been above sea level during periods of glaciation in the Quaternary, when the sea level was as much as 140 m below the current sea level. These lowstands have occurred at regular intervals over the past 500,000 years (Colonna et al. 1996, Rohling et al. 1998, Raxworthy et al. 2002). Although this period is too recent to account for the movement of large terrestrial cuckoos between southeast Asia and Madagascar and the subsequent speciation of couas and ground-cuckoos, the possibility of dynamic changes on ocean mounts and sea levels suggests an earlier but post-Gondwana opportunity for dispersal between the land masses. In the late Oligocene there were emergent ridges above sea level around the Mascarene Plateau and these ridges only subsided slowly thereafter. Phylogenetic resolution of the genetic relationships among other terrestrial vertebrates also supports a history of postGondwana dispersal between Madagascar and Southeast Asia (Raxworthy et al. 2002, Shapiro et al. 2002).

88 The Cuckoos The large subfamily Cuculinae includes more than half of all cuckoo species and occurs in both the Old and New World. The phylogeny indicates an origin in the Old World followed by a single invasion of the New World (Figure 5.1).The New World taxa comprise a single derived clade (Coccycua, Piaya, and Coccyzus), whereas the basal lineages within the group (e.g., Raffles’s Malkoha Rhinortha chlorophaea, The yellowbills Ceuthmochares) and the sister clade to Cuculinae (Couinae plus Centropodinae) are all Old World birds. Cuculinae also includes both nesting cuckoos and brood parasitic cuckoos. One of the most significant results of the mitochondrial genetic analysis is that brood-parasitism evolved in this clade on at least two separate occasions, once in the lineage leading to Clamator and once in the common ancestor of all other Old World brood parasites, tribe Cuculini (Figure 5.1). Three tribes are recognized: Rhinorthini, Phaenicophaeini (Figure 5.5) and Cuculini (Figure 5.6). A surprising result of the molecular analysis was the position of Raffles’s Malkoha Rhinortha chlorophaea, basal to the all other Cuculinae, including both nesting and brood-parasitic cuckoos. Rhinortha is highly divergent genetically and its placement as the sister group to all other taxa in Cuculinae is strongly supported. This result was obtained in separate analyses of the two different mitochondrial gene regions sequenced, and it was confirmed by sequencing three specimens of this species, all of which had nearly identical sequences. Rhinortha is the only extant member of a divergent clade recognized here as a tribe, Rhinorthini. The tribe Phaenicophaeini includes all other nesting species of Old World and New World Cuculinae as well as the brood-parasitic Clamator (Figure 5.5). This arrangement suggests that Clamator and the New World nesting cuckoos (Coccycua, Piaya and Coccyzus) were derived from malkoha-like ancestors. The large malkoha genus “Phaenicophaeus” in the broad sense of Delacour and Mayr (1945) and Schlegel (1862) is not a monophyletic group but represents a paraphyletic assemblage that includes the basal lineages in this tribe. The African yellowbill genus Ceuthmochares (in African tradition called a “coucal” by Layard 1874, Sclater 1930, Bannerman 1933, Roberts 1940 and Maclean 1993), a malkoha in its morphology,

chattered vocalizations and nesting behavior, is the sister group to all other species in Phaenicophaeini. The other malkohas (Taccocua, Zanclostomus, Phaenicophaeus, Dasylophus and Rhamphococcyx) may or may not form a monophyletic group: a clade comprising all malkoha species in these five genera was found in 2 of the 10 best trees from the weighted analysis on 171 taxa but not in analyses of Phaenicophaeini taxa only.The relative positions of Rhampococcyx, Dasylophus, and Clamator, in particular, were variable among analyses. Malkohas are diverse in morphology and plumage ornaments, and for many years these birds were classified in several separate genera, based on differences in the feathering on the head and the shape of the bill and nostril (Sharpe 1873, Shelley 1891). This view changed when Delacour and Mayr (1945) combined all the Asian malkohas and the African Ceuthmochares (separate genera that they were “unable to admit”) into a single genus Phaenicophaeus, with the assertion that “in general characters and habits the various Malcohas are so closely related that it seems more logical to consider them all as species of a single genus (Phoenicophaeus—1815).”The authors described no characters, even though several skeletal and other morphological differences among these cuckoos have been described by others (Sharpe 1873, Pycraft 1903, Seibel 1988, Hughes 2000). Delacour and Mayr apparently based their opinion on similarities they saw in skins. The genetic results (Figure 5.5) suggest the recognition of five Asian genera in addition to Rhinortha and the African genus Ceuthmochares. In Figure 5.5, six malkoha species are included in the genus Phaenicophaeus. Chestnut-breasted Malkoha P. curvirostris (previously included in Rhamphococcyx) is the sister group to a clade of more closely related species, Chestnut-bellied Malkoha P. sumatranus, Redfaced Malkoha P. pyrrhocephalus, Blue-faced Malkoha P. viridirostris, Black-bellied Malkoha P. diardi, and Green-billed Malkoha P. tristis (all but pyrrhocephalus formerly included in Rhopodytes).These birds have a large area of brightly colored bare skin around the eyes and sides of the face, an arched or deep bill, and many have contrasting colors on the bill. Relationships within Phaenicophaeus were generally stable among analyses, yet the result was not strongly

A molecular genetic analysis 89 Ceuthmochares aereus Ceuthmochares australis Taccocua leschenaultii 99 Zanclostomus javanicus (Java) Zanclostomus javanicus (captive) 80 Phaenicophaeus curvirostris 91 Phaenicophaeus sumatranus 93 Phaenicophaeus pyrrhocephalus 48 Phaenicophaeus viridirostris 70 Phaenicophaeus d. diardi 69 99 Phaenicophaeus tristis "tristis" Phaenicophaeus t. "longicaudatus" Rhamphococcyx calyorhynchus (2) 52 Dasylophus superciliosus 99 Dasylophus cumingi (2) Clamator coromandus 100 63 Clamator glandarius 95 Clamator levaillantii 100 Clamator jacobinus pica Clamator jacobinus serratus 8 Coccycua minuta 96 100 Coccycua pumila Coccycua cinerea (2) 100 87 Piaya cayana Piaya melanogaster Coccyzus melacoryphus 70 100 100 Coccyzus americanus Coccyzus euleri 97 100 Coccyzus minor 100 Coccyzus ferrugineus Coccyzus erythropthalmus 84 Coccyzus lansbergi 100 Coccyzus pluvialis (2) 36 95 Coccyzus rufigularis Coccyzus vetula 89 Coccyzus vieilloti Coccyzus merlini 58 100 50 changes Coccyzus longirostris 100

87

Figure 5.5. Phylogeny of the Old World “malkohas” and allies (tribe Phaenicophaeini).The tree shown is the majority rule consensus from the overall analysis, and is identical to the weighted analysis for Phaenicophaeini only.The position of Rhinortha chlorophaea is indicated in Figure 5.1. Alternative arrangements for Rhamphococcyx, Dasylophus, and Clamater are described in the text. Bootstrap values are shown above each node.Two samples of R. calyorhynchus, D. cumingi, C. cinerea, and C. pluvialis, respectively, had nearly identical sequences.

supported inasmuch as complete sequence data were obtained only for P. curvirostris. Two species, Sirkeer Malkoha Taccocua leschenaultii and Red-billed Malkoha Zanclostomus javanicus, are basal to Phaenicophaeus and are recognized as separate genera; they differ from other malkohas in the feathered face and straight bill with a single color. Yellow-billed Malkoha Rhamphococcyx calyorhynchus of Sulawesi and the crested malkohas Dasylophus of the Philippines generally did not group with other malkoha genera in our analyses, but instead branch off singly or together either before or after Clamator in most trees. In other trees, Rhamphococcyx and Dasylophus form a clade sister to Clamator. Although its placement is uncertain,

results for Rhamphococcyx are nonetheless surprising. Most systematic arrangements that recognize several genera of malkohas treat R. calyorhynchus and Chestnut-breasted Malkoha P. curvirostris as sister species in a single genus Rhamphococcyx. These two are morphologically similar and their geographic ranges are complementary. Whereas P. curvirostris is well known as a nesting cuckoo, only one nest of R. calyorhynchus has been described (Meyer 1879) and a fledgling was seen to be fed repeatedly by another species (Rozendaal and Dekker 1989). The fledgling malkoha was not necessarily reared in the nest of that species, insofar as songbirds can be attracted to a begging fledgling of another species and feed it repeatedly, especially when their own

90 The Cuckoos 20

Pachycoccyx audeberti Microdynamis parva Eudynamys s. scolopacea 100 Eudynamys s. melanorhyncha Eudynamys s. alberti Eudynamys s. cyanocephala (2) Urodynamis taitensis Scythrops novaehollandiae 100 Chysococcyx maculatus Chysococcyx xanthorhynchus (2) 76 Chysococcyx caprius (2) 100 Chysococcyx klaas 44 Chysococcyx flavigularis 100 Chysococcyx cupreus Chysococcyx cupreus Chysococcyx megarhynchus 85 Chysococcyx basalis 48 Chysococcyx osculans 62 18 Chysococcyx ruficollis Chysococcyx lucidus layardi Chysococcyx lucidus plagosus Chysococcyx meyeri (2) 14 Chysococcyx minutillus crassirostris 67 Chysococcyx minutillus poecilurus Chysococcyx minutillus barnardi 83 Cacomantis pallidus Cacomantis leucolophus 94 100 Cacomantis castaneiventris Cacomantis flabelliformis 73 Cacomantis sonneratii 67 Cacomantis merulinus 100 Cacomantis passerinus 100 Cacomantis variolosus “heinrichi” Cacomantis variolosus sepulcralis (3) Cacomantis variolosus variolosus Cercococcyx mechowii (2) 100 100 Cercococcyx olivinus 50 Cercococcyx montanus Surniculus dicruroides (2) 100 Surniculus velutinus (2) 95 Surniculus lugubris (2) Surniculus musschenbroeki 87 Hierococcyx vagans 96 Hierococcyx bocki 100 Hierococcyx sparverioides 96 Hierococcyx varius (2) 98 Hierococcyx hyperythrus 100 Hierococcyx pectoralis (2) 69 Hierococcyx fugax 100 Hierococcyx nisicolor (3) Cuculus c. clamosus Cuculus c. gabonensis (3) 93 Cuculus solitarius (2) 100 Cuculus poliocephalus (2) 88 Cuculus crassirostris Cuculus micropterus 82 Cuculus rochii 63 Cuculus gularis (2) 96 Cuculus saturatus (2) 100 Cuculus canorus Cuculus optatus (4) Cuculus saturatus 98 Cuculus canorus (5)

100

64 54

99 61

61

100

92

50 changes

Figure 5.6. Phylogeny of the Old World parasitic cuckoos (tribe Cuculini). Relationships within this group were identical in all analyses except for variations within Chrysococcyx, which are discussed in the text. A number following the taxon name indicates that multiple individuals were sequenced, and in general these individuals had identical or nearly identical sequences. Each of twelve C. canorus and C. saturatus samples had unique but similar sequences (four of which are shown here). In addition, the sequences from these two species did not form monophyletic groups (see chapter 6 species). Bootstrap values are shown above each node.

brood has recently failed and the birds are in the hormonal condition for parental care (Welty and Baptista 1988, Payne and Payne 2002). Nonetheless, this observation and the lack of information on the

nests of Dasylophus species are intriguing, particularly if Rhamphococcyx and Dasylophus form the sister group to Clamator. The reproductive biology of Rhamphococcyx and Dasylophus need to be

A molecular genetic analysis 91 investigated to determine if they are facultative or perhaps even obligate brood parasites. Perhaps just as important, the malkohas in these two genera are all east of mainland Asia and the sunda Shelf, in an area never connected to the Asian mainland (Lee and Lawver 1995, Hall, 1996), in the area originally described as the zoogeographic region east of Wallace’s line (Huxley 1868, Mayr 1994). The arboreal New World cuckoos Coccycua, Piaya and Coccyzus are closely related to the Old World malkohas and form a monophyletic group within the tribe Phaenicophaeini (Figure 5.5).This is in contrast to the account of Payne (1997b) which was based on an analysis of feather keratins (Brush and Witt 1983) suggesting a close relationship between Coccyzus and New World anis. Berger (1952, 1960) also questioned the idea that the coccyzine cuckoos were closely related to the malkohas, insofar as the two groups had a different pelvic muscle formula (AXYAm, vs ABXYAm). Sibley and Ahlquist (1990) grouped malkohas, Old World ground-cuckoos and couas with the Old World “Cuculidae” and apart from their New World family “Coccyzidae;” in fact none of these groups were sampled in their study.The mitochondrial evidence for including the arboreal New World cuckoos within Phaenicophaeini is strong and we refer to this clade as the coccyzine cuckoos. Within the coccyzine cuckoos the basal lineage is Coccycua. Little Cuckoo Coccycua minuta was called a Coccyzus by Vieillot in 1817 and a Piaya by Gray in 1846 and Shelley in 1891, and was described as a distinct genus Coccycua by Lesson in 1830. In the phylogeny (Figure 5.5) Coccycua includes two other species, C. pumila and C. cinerea, that were formerly listed as species of Coccyzus.Although C. minuta is like the Squirrel Cuckoos Piaya cayana and P. melanogaster in plumage color and pattern and in the long graduated tail, minuta is similar to Dwarf Cuckoo Coccycua pumila in the short, curved bill and rufous plumage (the rufous is restricted in pumila to the throat and breast), and pumila in turn is like Ash-colored Cuckoo C. cinerea in the shape of the bill and the rounded (not graduated) tail. C. cinerea was described as Coccyzus by Vieillot and as Piaya by Gray (Cabanis and Heine 1862). Although Dwarf Cuckoo C. pumila was more similar to C. cinerea than to other Coccyzus cuckoos in mtDNA cytochrome b sequence, Hughes (1997d ) compared only the cuckoos listed by Peters (1940) in

the genus Coccyzus, and did not include Coccycua minuta.When all these cuckoos are considered together, the molecular genetic data shows that the closest relatives of C. minuta are C. pumila and C. cinerea. The species pumila and cinerea were treated as the genus Micrococcyx by Ridgway (1912, 1916) based on their more rounded wing than that of other small New World cuckoos, and the genera Coccycua and Micrococcyx were considered to be separate based on a more rounded wing in minuta than in pumila and cinerea (Ridgway 1916). Shelley (1891) listed Coccyzus with Cuculus because they had long straight wings, and Piaya with Phaenicophaeus and other Old World malkohas because they had short curved wings. In Shelley’s view, minuta was a Piaya because it had short curved wings, while pumila was a Coccyzus because it had long straight wings.The molecular genetic results indicate that wing shape is not the most useful guide to the evolutionary relationships among these New World cuckoos. The large Squirrel Cuckoo Piaya cayana and Black-bellied Cuckoo P. melanogaster are each others’ closest relatives and form the next clade within the coccyzine cuckoos. The other New World coccyzine cuckoos are recognized here in a single genus Coccyzus. Most species are small and have long pointed wings. South American Pearly-breasted Cuckoo C. euleri and North American Yellow-billed Cuckoo C. americanus are sister species which are in turn sister to the closely related Mangrove Cuckoo C. minor and Cocos Cuckoo C. ferrugineus.The large cuckoos of the West Indies that have generally been referred to as Hyetornis, and the lizard-cuckoos Saurothera, appear to be derived from smaller cuckoos within the genus Coccyzus.The genus Coccyzus is monophyletic only if these large island cuckoos are included, and this is the basis for combining all these birds into a single genus. The genetic, morphological and distributional data suggest that these large cuckoos of the West Indies were derived from the continental Coccyzus cuckoos, perhaps birds like South American Gray-capped Cuckoo C. lansbergi, which some of them resemble in plumage color and pattern.The large island cuckoos are also similar to continental Coccyzus in the sound of their throaty and guttural rattling calls.The cuckoos of the West Indies evolved large body size and rounded wings (P8 is one of the shorter primaries)

92 The Cuckoos on the islands, a trend that other kinds of birds have undergone as well (e.g., Livezey 1993, 1998). Evolution of large body size of island cuckoos is also indicated in Mangrove Cuckoo C. minor, which is larger than its close relatives C. americanus and C. euleri, and in Cocos Cuckoo C. ferrugineus, an isolated island cuckoo that is larger than its allospecies, C. minor. Within the West Indies, the Hyetornis and Saurothera species groups apparently evolved by repeated invasions across the Greater Antilles, where two islands (Hispaniola and Jamaica) each have two species, one from each group. In the phylogeny of Clamator, C. coromandus rather than C. glandarius is basal to the other species. This result is inconsistent with a division of these cuckoos into two genera, Clamator and “Oxylophus” ( jacobinus and levaillantii ) (Wolters 1975–1982, Sibley and Monroe 1990).Those studies apparently considered the Asian species C. glandarius to be more similar to C. coromandus than to the two black-and-white species, but they did not consider discrete characters or the evolutionary branching sequence among these cuckoos. Perhaps the most significant result of the molecular phylogeny is the implied double origin of brood parasitism in the Old World cuckoos, a result also found by Aragón et al. (1999). The crested cuckoos Clamator are more closely related to the malkohas and coccyzine cuckoos in Phaenicophaeini than to other Old World brood parasites, indicating that these crested cuckoos evolved obligate brood parasitism independently of the other Old World brood parasites.This conclusion is supported by several anatomical and behavioral traits. (1) Brain size relative to body weight in Clamator is like that of the nesting cuckoos, and unlike the small brains of the other Old World brood-parasitic cuckoos. (2) Both Clamator and other malkoha genera have elongate rather than raised circular nostrils, and both have the same formula of thigh muscles, AEXYAm; in Phaenicophaeus pyrrhocephalus and Ceuthmochares an additional muscle is present (Berger 1960).The same groups have a double-notched sternum, unlike most other Old World brood parasites in Cuculini; however, several other anatomical characters that vary among cuckoo genera do not sort out along these same lines (Berger 1960, Seibel 1988, Hughes 1997d, 2000). (3) Egg size in relation to body weight is larger

in Clamator and the malkohas than in the other Old World brood-parasitic cuckoos, which lay eggs that are relatively small for their body size. (4) Clamator have a plumage crest, like the closely related Dasylophus malkohas and unlike the other Old World brood-parasitic cuckoos. (5) Clamator and the lizardcuckoos Coccyzus (Saurothera) share an unusual molt pattern in which wing molt progresses in a regular pattern that skips two feathers, with a sequence P6 9 - pause - 7 - 10 - pause - 8 - 5 (or - 5 - 8). (6) The calls and songs of Clamator are harsh repeated sounds, like the calls of malkohas but unlike the whistled songs of most Old World brood-parasitic cuckoos. (7) Clamator resemble the nesting Old World and New World malkohas in having a pair bond; the male assists in brood parasitism by accompanying the female to host nests. (8) Nestling Clamator like nestling malkohas do not evict eggs from the nest, as do other Old World brood parasites. (9) The begging posture of Clamator nestlings is unlike that of nestling brood-parasitic Cuculus and Chrysococcyx, which crouch in the nest and hold their head low over the back and do not flutter their wings, and do so only inconspicuously after they fledge. Young Great Spotted Cuckoos Clamator glandarius stretch their neck upwards in high-intensity begging, much as do nesting Black-billed Cuckoos Coccyzus erythropthalmus, Chestnut-bellied Malkohas Phaenicophaeus sumatranus, Red-capped Couas Coua ruficeps, and Coquerel’s Couas Coua coquereli (Appert 1970, 1980, Ong Kiem Sian, Figure 5.7). At other times they crouch in the nest (von Frisch and von Frisch 1967, M. Soler, in litt.). Begging postures of young Jacobin Cuckoos Clamator jacobinus also are upright with the neck stretched upwards at an early age (Liversidge 1971), and fledged Levaillant’s Cuckoos C. levaillantii conspicuously flutter their wings (Vincent 1934), unlike Cuculus and Chrysococcyx cuckoos. However, when nestling Rufous Hawk-cuckoos Hierococcyx hyperythrus beg, they raise the wings in display to their foster parents. (10) Young nesting cuckoos grow rapidly and fledge early, whereas brood-parasitic cuckoos grow slowly and let the parents provide prolonged parental care in the nest. Nestling growth in young C. jacobinus is slower than in nesting cuckoos such as Black-billed Cuckoo and Yellow-billed Cuckoo, but is faster than in Cuculus of the same adult body size. Nestling growth in Clamator jacobinus, for example, is

A molecular genetic analysis 93

Figure 5.7. Begging behavior of nestling cuckoos: a, Groove-billed Anis, Costa Rica; b, Black-billed Cuckoo, Michigan; c, Horsfield’s Bronze-cuckoo,Western Australia.

more rapid than in Cuculus solitarius, with Clamator reaching half of fledging weight (60 g) by day 7 rather than day 9. Like nesting cukoos, C. jacobinus and C. levaillantii also have shorter nestling periods (11–14 days) than other broodparasitic cuckoos of the same body size. Comparison with the Old World malkohas is not possible inasmuch as the details of their breeding biology are unknown. Old World brood-parasitic cuckoos, except for the crested cuckoos Clamator, share a common ancestor and are each others’ closest relatives in most of the best trees from our overall analyses.These Old World cuckoos are recognized as a tribe Cuculini (Figure 5.6). A sister relationship between the Thick-billed Cuckoo Pachycoccyx audeberti and the

koels (Eudynamys and Microdynamis) was consistently found in our analayses, but was weakly supported by the measure of bootstrap support.This clade was the sister group of all other Cuculini in most trees. In 2 of the 21 best trees from the overall analysis, however, the clade comprising Pachycoccyx, Microdynamis, and Eudynamys is basal to both Phaenicophaeini and other Cuculini, branching off just after Rhinortha, a result that would imply a fourth origin of brood parasitism in cuckoos.Additional data are needed to further test the monophyly of Cuculini. Channel-billed Cuckoo Scythrops novaehollandiae and Long-tailed Cuckoo Urodynamis taitensis of Australasia are each other’s closest relatives and form the next most basal lineage within Cuculini.

94 The Cuckoos The glossy cuckoos Chrysococcyx are monophyletic (Figure 5.6). The African and Australasian species form two well-supported clades and the Asian C. maculatus and C. xanthorhynchus appear to be more closely related to the African glossy cuckoos than to the Australian and New Guinea bronze-cuckoos. In some analyses, however, the two Asian species are the basal lineage within the genus. Long-billed Cuckoo C. megarhynchus, once recognized as a genus Rhamphomantis by itself, is a member of the Australian and New Guinea glossy cuckoo lineage within Chrysococcyx. The placements of C. basalis and C. oscualans were variable among analyses, with C. basalis sister to C. megarhynchus and/or C. osculans sister to the clade comprising C. meyeri and C. minutillus in some trees. Incomplete data for C. megarhynchus and C. ruficollis likely contribute to this uncertainty. Other brood-parasitic cuckoos with gray and brown plumage and barred underparts form a monophyletic lineage (Figure 5.6).This group includes the brush cuckoos Cacomantis, a clade comprising the African long-tailed cuckoos Cercococcyx and Asian drongo-cuckoos Surniculus, and a clade comprising the hawk-cuckoos Hierococcyx and the canonical cuckoos Cuculus. Realtionships within this entire clade were generally well-supported and no variations in tree topology were found in our analyses.The Hawk-cuckoos Hierococcyx are the sister group of Cuculus cuckoos, but not all birds that have been called “hawk-cuckoos” are each others’ closest relatives. Sulawesi Cuckoo Cuculus crassirostris, which was known as a “hawk-cuckoo”, is not closely related to either the H. fugax or H. sparverioides species complexes, but is more closely related to the more typical (non-hawk) Cuculus cuckoos such as C. micropterus and C. canorus. C. crassirostris was originally described as a Hierococcyx and it was long called a “hawkcuckoo” because it has broad, rounded wings and a broadly barred tail (Meyer 1879, Shelley 1891, Baker 1927, Stresemann 1940, Wolters 1975–1982), but its calls are mellow whistles like the calls of Common Cuckoo C. canorus, and its plumage below is barred not streaked; it is better known as Sulawesi Cuckoo. Wolters (1975–1982) listed Indian Cuckoo C. micropterus in his genus Hierococcyx based on its broadly barred tail, but C. micropterus is also in the Cuculus lineage, both in the molecular genetics’ results

and in its rhythmic whistled song. Drongo-cuckoos Surniculus and White-crowned Cuckoo Cacomantis (Caliechthrus) leucolophus were once thought to be related to koels Eudynamys (Peters 1940), if only because the adult plumage is black (as in adult male Common Koel). The genetic results indicate that both drongo-cuckoos and White-crowned Cuckoo are more closely related to Hierococcyx and Cuculus than to the koels. Drongo-cuckoos and Whitecrowned Cuckoo resemble Cuculus and Cacomantis in the step-graded whistles of their song (especially Fork-tailed Drongo-cuckoo S. dicruroides and Pallid Cuckoo Cacomantis pallidus), and a bubbling crescendo call (as in Black Cuckoo Cuculus clamosus and in hawk-cuckoos of the Hierococcyx fugax complex). The mitochondrial sequence data presented above provides a remarkably good resolution of the evolutionary relationships among cuckoo subfamilies, genera and species, and a framework for future analyses of cuckoo systematics and population biology. Nonetheless, our results should be considered preliminary, in that additional data and more thorough analyses of the currently available data are needed. Sequence data from the nuclear genome needs to be collected to test our conclusions and resolve some of the uncertainties noted above. Issues in need of further analysis include the relationships among cuckoo subfamilies; monophyly of Cuculini and the position of Pachycoccyx; and the relationships among Rhamphococcyx, Dasylophus, Clamator and other malkohas. At the species level, more data are needed for Carpococcyx viridis, Coua delalandei, some of the coucals Centropus, various malkohas Phaenicophaeus, and Chrysococcyx megarhynchus.A number of clades that were consistently found in our analyses were weakly supported by the measure of bootstrap value (Figures 5.1–5.6), and it is these clades that may be most likely to change as additional data become available. Finally, there are many opportunities for population level studies exploring geographic variation and species limits, particularly within the Old World brood parasites (Cuculini). The lack of reciprocal monophyly in the mitochondrial DNA of C. canorus, C. optatus and C. saturatus, for example, needs to be examined with a broad geographic sample of populations of these two species.

6 Species

What is a species? In his book On the origin of species, Darwin (1859, 1875) recognized a species as a morphologically distinct class of individuals. He avoided a formal definition of the term, and he even referred to “the vain search for the undiscovered and undiscoverable essence of the term species”. Earlier, when Linnaeus (1758) wrote Systema naturae and established systematic biology, he described species as morphologically distinct kinds.A more current idea recognizes a biological species as a class of individuals that interbreed with each other, at least when they live in the same population (Mayr 1963, Klein and Takahata 2002). In most kinds of birds it is clear whether individuals are members of the same species, where sex for sex and age for age the birds look alike within a population and among geographic populations. However, some populations live in different regions and the birds can differ somewhat in plumage but otherwise appear to be the same kind of bird.This geographical separation of populations can be a challenge to the biological species concept (Cracraft 1983). However, most biologists find reasonable and rewarding the notion of species being composed of populations with restricted gene flow between these and other populations (Avise 1994). In birds, if pairs that feed their young at the nest are mated, exclusively from birds that look different, they can be considered to be behaving as a distinct biological species. However, the brood-parasitic cuckoos do not live together in pairs near the nest as they have no nest. Males and females do not generally associate together and are not seen as pairs

except briefly when they copulate, or more often when the male feeds the female before they mate. In most brood-parasitic cuckoos it is impossible to observe mated pairs, and only when lucky we see the social associations that lead to mating. Only in the few brood-parasitic cuckoos where male and female visit a nest together (as do the crested cuckoos, especially Jacobin Cuckoo Clamator jacobinus, where the pair visits the nest of a host), do mated pairs live together. In southern Natal, where two color phases occur together, many pairs are assortative, the black-phase birds mated with black-phase birds and the white-bellied birds mated with white-bellied birds. This suggests some degree of like mating with like. Nevertheless, the songs of black-phase and white-phase birds in this region are the same. Also, further south in the Eastern Cape where white-phase birds are outnumbered by black-phase birds, all white-phase birds observed are mated with black-phase birds. The similarity of songs and the common mating between the two color phases indicates that these Jacobin Cuckoos are members of a single species. Perhaps most useful in identifying species of cuckoos is their song. Of course the birds do not sing to us, they sing to other birds, and they use song to attract a mate. Among cuckoos, the birds that share their songs and morphology are recognized as biological species. In the field we can test whether birds respond to playback of recorded song equally well to foreign songs and to local songs, and if they do so respond, this equal responsiveness indicates that they are the same species (Payne 1986, Alström and Ranft 2003). In addition to their appearance and song, another criterion of species distinctiveness

96 The Cuckoos is whether two forms live together without interbreeding: the test of sympatry. If they live together and do not interbreed, then the populations are distinct species (Mayr 1963). Although certain birds occasionally hybridize even when they are not each others’ closest rela-tives (e.g., orioles Icterus: Freeman and Zink 1995, Omland et al. 1999), interspecific hybrids are unknown in the cuckoos. Genetic information is the best source of data to estimate the phylogenetic relationships among species. But species pairs of birds differ from each other in degrees of genetic divergence. That is, we have no molecular genetic criteria that define a species or tell us which allopatric populations that differ in appearance, are distinct enough to recognize as species. In the molecular genetic comparisons (Figures 5.5, 5.6), some species pairs differ in genetic distance by several times that of other species pairs (perhaps some species in one lineage became extinct, so the lineages did not diverge at the same time and not all species pairs at present are equally old), and some cuckoos are more variable within a species (e.g. Zanclostomus javanicus, Chrysococcyx cupreus) than in some cases between species.The genetic distance between the species Cacomantis merulinus and C. passerinus is less than between populations within some species such as Chrysococcyx cupreus, but the songs of the first two cuckoos are distinct, whereas the songs and appearance of birds in populations of C. cupreus are all similar and they are regarded as a single species. Genetic information can tell us which kinds of birds are each others’ closest relatives and share a common ancestor. Genetic monophyly is a condition necessary to define a clade, but monophyly is not necessary or sufficient to define a species. Generally we expect members of a species to form a monophyletic class, but then also we expect the same for members of a clade at any higher systematic level such as a genus. In addition, gene monophyly is not a necessary criterion of a species, insofar as gene trees do not always coincide with species trees. In some cases, two species live together and do not interbreed, yet they do not have mutually exclusive sets of genes. Genes can be shared between recently diverged species when the genes in their ancestral population have not yet become differentially

extinct in each daughter species. In consequence, the evolutionary divergence of gene trees can lag behind that of species trees (Avise 1994, Klein and Takahata 2002).A case of this was found in the cuckoos, where Oriental Cuckoos Cuculus optatus. Himalayan Cuckoos C. saturatus, and common Cuckoos C. canorus do not consistently differ in mtDNA gene sequences (Figure 5.6). The overlap of genetic sequence between these two species may result from recent speciation and incomplete lineage sorting or unilateral extinction of genetic morphs that were present in their common ancestor. Lack of reciprocal genetic monophyly might also result from hybridization, but we know of no case of hybridization in the cuckoos (as mentioned in the species account for Yellow-billed Cuckoo). In these Cueulus species the birds differ morphologically in their songs, so they are distinct species. Except in this one case, mtDNA monophyly was the rule among cuckoo species that were compared in more than one individual. When populations differ consistently in morphology and in song, they are distinct species: together these are sufficient criteria to recognize species. On the other hand, there is no general standard measurement of genetic difference to say that populations are one species or more than one species. For this reason, morphological and biological criteria are more general criteria to use in recognizing a species, than are degrees of genetic differences between populations. With this rationale it was possible to use two criteria to recognize species as distinct. First, if two or more kinds of birds live together and do not interbreed, then they are considered distinct species. Second, if populations have the same calls and songs, they are considered one species, and if they have different calls, they are different species.

Songs and species Birds use songs in choice of mate, and males use songs in behavior towards competitors, chasing away other birds with the same song and excluding them from their territories, while ignoring birds with different songs. The response of males in attacking a tape recording of songs of their own species while ignoring songs of other kinds of birds is often used as a test of whether birds are distinct

Species 97 enough to be considered different species (Payne 1986). Females are more difficult to test because females often do not respond to song playback in the field, but when females do respond, they are as selective in response as the males (Searcy 1992). Female cuckoos have not been tested for their choice of song, and our use of song to indicate species identity is based on what is known about birds that do respond to song in this way, as well as on the response of male cuckoos to song.The reasoning is that females use song to choose a mate and the specificity of male song determines the set of potential mates. Among cuckoos it is reasonable to use song in the same way as the birds themselves, in recognizing a species when they hear the song. The songs and calls are remarkably similar through the range of a cuckoo species (Chappuis 1974, 2000).The Common Cuckoo Cuculus canorus “cúck-oo” is a standard through the Palearctic from Britain to China and Japan (allowing for songs out of tune when the breeding season ends). Songs of the Oriental Cuckoo C. optatus are also the same across a continuous expanse in northern Asia. In Africa the songs of the African Cuckoo sound the same from Gambia and Nigeria through Zambia and South Africa. Similarly, songs of Diederik Cuckoo Chrysococcyx caprius, Klaas’s Cuckoo C. klaas and African Emerald Cuckoo C. cupreus sound the same across the range of each species. From Western Australia to New Zealand, Shining Bronze-cuckoos C. lucidus are alike in their songs. The songs also sound the same to the cuckoos. Laura and I compared the response of cuckoos in Western Australia to the playback of cuckoo songs recorded in New Zealand, to songs of cuckoos in their own Australian population, and to songs of another species of cuckoo.The cuckoos did not respond to songs of Horsfield’s Bronze-cuckoo C. basalis, but they responded by approaching, calling and flying over the speaker to songs of Shining Bronzecuckoo from New Zealand as quickly as they did to songs of their own population. Although New Zealand cuckoos and Western Australian cuckoos differ in plumage with females in New Zealand looking like males in Australia, the birds are the same species on the basis of performance and response to their common songs.

The similarity in song across a wide geographic range of these cuckoos suggests that cuckoo song is innate and not learned. A behavior is said to be “innate” when a bird develops the normal behavior whether or not it has any contact with another individual of its own species (Lorenz 1935). Indirectly this term is sometimes taken to indicate a “genetic” basis. Mayr (1942) argued,“There is also no question that the song of parasitic cuckoos must have a strictly genetic basis, for the young cuckoo is obviously more exposed to the song of his foster father than to that of his own”. However, the term “innate” refers to the social conditions of development, and not to genetics. There is some evidence for innate song development in cuckoos. A young Channel-billed Cuckoo was reared in captivity and developed normal adult calls (Goddard and Marchant 1983). Species songs are nearly the same across a wide range of nesting cuckoos as well, as they are in Yellowbilled Cuckoos Coccyzus americanus across North America. Call development may be innate in Greater Roadrunner Geococcyx californianus, as hand-reared birds that do not hear their own species have adult calls like the calls in wild adults (Whitson 1971), and a hand-reared Smooth-billed Ani Crotophaga ani developed normal calls like those of wild birds (Quinn and Startek-Foote 2000). Because their songs are similar across a species range of cuckoos, we use song as a guide to the species of the less well-known cuckoos to determine the biological limits of species. If songs are the same in geographically separated populations, then the birds are likely to be conspecific; and if songs are different (allowing for song repertoires), then the populations are likely to be two or more species. If the birds differ both in morphology and in song, then the birds are distinct species.This guideline has led to the discovery and recognition of several new species of birds by their songs in recent years (Alström and Ranft 2003). Using their song as a guide, African Cuckoo Cuculus gularis is a species distinct from Common Cuckoo C. canorus.African birds sing with an accent on the second note of “cuckóo,” and this note rises rather than drops in pitch from the first note. Palearctic birds sing with an accent on the first, higherpitched note; the familiar “cúck-oo.” In playback

98 The Cuckoos tests in the field, males were attracted to their own kind of song, but not to the song of the other species (Payne 1986). In addition, the birds differ in morphology—in adult and juvenile plumage, in bill color (yellow in African Cuckoo, black in Common Cuckoo) and in bill width, wide at the base in African Cuckoo and narrow in Common Cuckoo (Payne 1977a).These cuckoos were indifferent to the songs of the other kind of cuckoo, and their songs are different.We use this difference in their songs as a standard to recognize other species of cuckoos. Song recordings have identified more species of cuckoos in the Old World than was done before their songs were studied. In a few cases, their distinctive songs were identified only in the past few years. (1) The number of hawk-cuckoo species recognized in the Hierococcyx fugax complex has increased from one (Salvadori 1874, Shelley 1891, Peters 1940, King and Dickinson 1975) to two (Payne 1997b) or three (King 2002), or even four.These hawk-cuckoos were described as separate species in the early 19th century, the latest one by Gould (1856) at a time when populations with any morphological differences were described and published as separate species.These hawk-cuckoos were then most dramatically combined into a single species by Salvadori (1874), who included not only nisicolor, hyperythrus and pectoralis with H. fugax but also sparverioides and varius within this species. He did not include H. vagans or its synonym H. nanus in his review, which he wrote before he worked at Leiden (where the holotype of C. vagans is kept) and at the British Museum (where Hume’s type series of H. nanus is kept) (Shelley 1891, Warren 1966,Violani et al. 1997). Shelley (1891) recognized H. varius, H. sparverioides and H. bocki as distinct species, but he combined the forms nisicolor, hyperythrus and pectoralis and even H. vagans with H. fugax.The larger forms were considered subspecies of H. fugax by Mayr (1938a), on the basis of being geographic counterparts. Only after their songs were recorded and found to co-vary consistently with the plumage forms were the four hawk-cuckoos Hierococcyx fugax, H. nisicolor, H. hyperythrus and H. pectoralis again recognized as distinct species. Songs of Philippine Hawk-cuckoo Hierococcyx pectoralis (“wheet wheet wheet . . .”) are distinct from songs of Javan Hawk-cuckoo H. fugax (a shrill “gee-

whizz, gee-whizz”) (Payne 1997b). Songs of Rufous Hawk-cuckoo H. hyperythrus in northern China, Russia and Japan are even more distinct, a loud harsh call “wee wee-pit” (Figure 6.1). The songs in Japan are so different from songs in the Himalayas that I suspected a recording problem and excluded the Japanese songs in an earlier review of cuckoo species (Payne 1997b), but P. Alström and B. King have recorded the same songs in northeastern China. Songs of Whistling Hawk-cuckoo H. nisicolor in the Himalayas are similar to songs of H. fugax in the Malay Peninsula and the Malay Archipelago, but there are differences. Southern birds H. fugax have lower-pitched notes than northern birds H. nisicolor, with the first note loudest between 3 and 4kHz in H. fugax and the first note loudest between 4 and 5kHz in H. nisicolor. In addition, in southern birds H. fugax the first note in the song rises in pitch through its length; northern birds H. nisicolor hold the first note on one pitch for most of its length before it rises. In the recordings available, six H. fugax in southern Thailand and the Malay Peninsula (NSA 15709, 32612, 34934, 41317; Scharringa 1999) have songs lower in pitch and with the first note sliding upward in pitch, compared with northern birds H. nisicolor (Sichuan, P.Alström; Guangdong, B. King; northern India, P. Singh; Bhutan, S. Connop 1995; northern Thailand, P. Round; Figure 6.1).Across the species complex, northern birds hold the first note on one pitch the longest time and Philippine birds hold it the shortest time. In addition to these songs, the four species of hawk-cuckoos give a long call that rises and then falls in pitch. Northern birds differ in wing shape as well as in song from southern birds. The outer primaries are longer than the inner primaries in the migratory cuckoos in northeast Asia (H. hyperythrus) and the Himalayas (H. nisicolor); the outer primaries are shorter and the wing is more rounded in the cuckoos of the Malay Peninsula and Greater Sunda Islands (H. fugax) and the Philippines (H. pectoralis).The species accounts describe other morphological differences. (2) The small southern Dark Hawk-cuckoo H. bocki and the large northern Large Hawk-cuckoo Hierococcyx sparverioides have different songs, perhaps more different from each other than either is from Common Hawk-cuckoo H. varius (Figure 6.2).

Species 99

Figure 6.1. Hierococcyx fugax complex: a–c, Rufous Hawk-cuckoo H. hyperythrus (a, Nagano, Japan, LNS 08259; b, Japan, LNS 08260; c, Heilongjiang, China, P.Alström); d–f,Whistling Hawk-cuckoo H. nisicolor (d, Sichuan, P.Alström; e, Bhutan, Connop 1995; f, Khao Yai NP,Thailand, P. Round); g–i, Javan Hawk-cuckoo H. fugax (g, Krabi,Thailand, Scharringa 1999; h, Ampang, Malay Peninsula, NSA 94048; i, Ampang, NSA 15709); j–k, Philippine Hawk-cuckoo H. pectoralis (j, Mt Katanglad, Mindanao, Philippines, NSA 44887; k, Baralatan, Mindanao, Philippines, NSA 34307).

Figure 6.2. Larger hawk-cuckoos: a, Common Hawk-cuckoo Hierococcyx varius (N Pakistan, P. Alström); b, Large Hawk-cuckoo H. sparverioides (Uttar Pradesh, India, P. Alström); c, Bock’s Hawk-cuckoo H. bocki (Malay Peninsula, White 1984).

Insofar as the two differ with no overlap in size, adult plumage color and juvenile plumage as well as in voice, they appear to be distinct species, as Shelley (1891) concluded from their morphological differences. The molecular phylogeny indicates that H. bocki is basal to these other two hawk-cuckoos, and that bocki and sparverioides are not a monophyletic set and so are not conspecific. (3) The northern and southern Asian drongocuckoos formerly were thought to comprise a single species, Surniculus lugubris, as in Peters (1940) and

Payne (1997b). Nevertheless, songs of Asian birds and Philippine birds were recognized as different by several field observers, and their plumage differs as well. The juvenile plumage of the Philippine Drongocuckoos S. velutinus is black with white spots in the Asian birds, uniformly rufous in southern Philippine S. velutinus (although black with white spots in the northern Philippines) and, the adult plumage has a glossy mantle and forked or square tail in Asia, and a velvet mantle and square tail in the Philippines. Songs of drongo-cuckoos differ in the number of notes, in

100 The Cuckoos change in pitch within the notes, and between notes in a song. Songs in southern Asia are usually of six notes sung in arpeggio,“one, two, three, four, five, six” rising smoothly up the scale (Legge 1880, King and Dickinson 1975). Songs in the Philippines have 8 to 10 notes, given more rapidly in a series, and the notes are closer to one pitch (Figure 6.3). Moluccan Drongo-cuckoo S. musschenbroeki in the northern Moluccas and Sulawesi differ from Philippine birds in song with a long series of notes: each note drops in pitch in the Moluccas, whereas each note rises in Philippine birds. Finally, songs of many drongo-cuckoos in mainland Asia from the Himalayan region, eastern Assam (Dig Boi) and the Chin Hills of Burma to northern Vietnam and China do not progress smoothly up the scale. Rather, the second note is no higher than the first, especially at the beginning of the note (sometimes the end of the second note is higher than the end of the first). These birds are large and have a deeply forked tail and are Fork-tailed Drongocuckoos S. dicruroides. In contrast, drongo-cuckoos of western Assam, central Thailand, the Malay Peninsula, Sri Lanka, Sumatra, Borneo and Palawan have songs

that rise uniformly through the series of notes, and the individual notes are more mellow in tone, less strident and piping, than notes of Fork-tailed Drongo-cuckoos.These more lowland and southern birds are Square-tailed Drongo-cuckoos S. lugubris. The occurrence of both kinds of songs in Assam suggests sympatry of two species of drongo-cuckoos. And in Sri Lanka the presence of both square-tailed and fork-tailed birds during the breeding season suggests two species; only the songs of S. lugubris appear to have been recorded there. Birds were not recorded as they sang in the field and then captured or collected to confirm a match of song and morphology, but in the information available on songs and morphology (tail shape), it appears that there are two species of drongo-cuckoos in India and southeast Asia.The two drongo-cuckoos overlap in their breeding areas.The extent of geographic overlap and morphological differentiation of these two cuckoos that are best identified by their songs remains to be worked out. (4) Indian Gray-bellied Cuckoo and rustybellied races of Plaintive Cuckoo (Cacomantis

Figure 6.3. Drongo-cuckoos: a–f, Fork-tailed Drongo-cuckoo Surniculus dicruroides (a, Himachal Pradesh, P. Singh; b, Kathmandu Valley, Nepal, NSA 12449; c, Dig Boi oilfields, Assam, NSA 100506; d, Mt Victoria, Chin Hills, Burma, P. Alström; e, Hainan, P. Alström; f, Ho Ke Go, Vietnam, NSA 46844); g–m, Square-tailed Drongo-cuckoo S. lugubris (g, Kaziranga,Assam, NSA 15022; h, Sri Lanka, Deepal Warakagoda; i, Huai Kha Khaeng,Thailand; j,Ampang, Malay Peninsula, NSA 31616; k, Sumatra, NSA 06806; l, Sabah, NSA 36035; m, Palawan, NSA 17837); n–o, Philippine Drongo-cuckoo S. velutinus (n, Mindanao, NSA 65939; o, Luzon, NSA 16435); p–r, Moluccan Drongo-cuckoo S. musschenbroeki (p, Sulawesi, NSA 68079; q, Sulawesi, LNS 32990; r, Halmahera, NSA 32969).

Species 101 passerinus and C. merulinus) in southeast Asia, Malaysia, Indonesia and the Philippines differ in song. Indian birds C. passerinus give a “piteer” call for their rising song, while the southeast Asian birds C. merulinus give “eat more froueet”; and the two cuckoos also differ in the cadence call (Figure 6.4). Insofar as their songs and morphology both differ, these two sets of populations are now considered distinct species, in contrast to reports that were made before good samples of songs were compared (Payne 1997b). (5) Three small Cuculus that are similar in adult plumage now are recognized as distinct species because they differ in juvenile plumage, in song, and (at least in the two species that were examined) in molecular genetics that indicate each is more closely related to a different cuckoo species than to each other.The three previously were recognized as a single species, “Lesser Cuckoo” C. poliocephalus (Peters 1940).The characteristic plumages of adults and juveniles are described in the species accounts, and the phylogenetic relationships are illustrated in chapter 5. The songs are illustrated for the three species, Madagascar Lesser Cuckoo C. rochii, Asian Lesser Cuckoo C. poliocephalus and Sunda Lesser Cuckoo C. lepidus in Dowsett and Dowsett-Lemaire (1980),Wells (1982), Becking (1988) and King (2005), and the song differences between all three species first were pointed out by Dowsett and Dowsett-Lemaire (1993). In other cuckoo species, the songs are the same between populations that are geographic counterparts. This similarity in song indicates that the birds would respond behaviorally to each other. Several sets of geographically replacing forms of brood-parasitic cuckoos in the Old World have nearly identical songs.

(1) Indian Cuckoos Cuculus micropterus in northern and southern Asia have nearly identical songs except for a slightly lower pitch in the northern birds, which are often larger but overlap in size with the southern birds. (2) Common Koels Eudynamys scolopacea throughout Australasia have similar songs,The koels are sometimes split by systematists into two or more species (“E. melanorhyncha”, “E. orientalis” and “E. cyanocephala”) by bill color and female plumage. Among these cuckoos each bird has a repertoire with more than one call (“ko-el”,“ke-woo” or “who are you” and “kwow kwow-kwow-kwow” in Asia (E. s. scolopacea); “kuOw” or “whooo” and “oeoeoeoeoeoe” in Sulawesi (melanorhyncha); “cooee” or “kooeel” and “wurroo” in Australia (cyanocephala). Apparently these are two sets of calls given in different contexts and each have their counterparts in each region. More field recordings are needed to test whether the corresponding calls are the same. (3) Northern and southern populations of Brush Cuckoos Cacomantis variolosus (combining the graybreasted and the rusty-breasted or Indonesian cuckoos “C. sepulcralis” ) have a similar rising song (“where’s the tea?”) and cadence song (“fear-fear”). The songs are not identical in northern and southern birds, and the evidence for the birds being conspecific involves the morphological intergradation of populations that live in the Moluccas and New Guinea.The rising song is longer and lower in pitch in the larger southern birds, yet when the birds are excited they shift into a high pitch that is nearly identical in Australia with the rising song of birds in the Malay Peninsula and Sulawesi (Figure 6.5).The

Figure 6.4. Plaintive cuckoos Cacomantis passerinus and C. merulinus: a–d, Gray-bellied Cuckoo C. passerinus, (a, b, “kateer” (a, Himachal Pradesh, Pratap Singh; b, Kaziranga, Assam, NSA 14045); c, d, “pee-pipee-pee-pipee” (c, Thekkady, Kerala, NSA 15016; d, Manas, Assam; NSA 15017)); e–h, Plaintive Cuckoo C. merulinus: (e, f, “eat more froueet” (e, Mt Victoria, Chin Hills, Burma, Per Alström; f, Sulawesi, Smith 1993a); g, h, cadence song (g, Mt Victoria, Burma, Per Alström; h, Mindanao, Philippines, NSA 31714)).

102 The Cuckoos descending “fear” calls may differ between areas, and additional recordings are needed to compare the behavior of northern and southern birds and test whether they are conspecific. (4) In the Little Bronze-cuckoo Chrysococcyx minutillus geographic complex, two or more species are sometimes recognized (Parker 1981). C. minutillus occur in many geographic forms from southeast Asia through the Malay Archipelago and the Philippines to New Guinea and northern and eastern Australia.The songs of C. m. minutillus, C. m. poecilurus, C. m. crassirostris and other geographic populations appear to be identical (Figure 6.6). The plumages of C. m. minutillus and C. m. poecilurus intergrade where their ranges come together in northern Australia. Because their trilled songs are similar throughout their overall geographic range, the green-backed forms (minutillus and others) and rusty-backed forms ( poecilurus and others) appear to be a single species, and the morphologically distinct pied form crassirostris has the same trilled song. None of the forms are known to occur together in their breeding range, and song as well as morphologically intermediate individual birds are sufficient criteria to consider these birds as a single biological species.

(5) African Black Cuckoos Cuculus clamosus differ remarkably in appearance across their range in Africa. Birds breeding in southern Africa are nearly completely black in plumage, and birds in the forests of central Africa are rufous on the throat and are barred and appear more or less white on the underparts. These birds have the same “papa’s boy” slurred whistles for their song, and birds in intermediate plumages occur where the extreme color forms come into geographic contact, so these cuckoos are considered a single species both on the basis of their song and on the basis of the morphologically intermediate individuals, even though they are difficult to spot while mating and there is no observational evidence of whether mated cuckoos tend to be similar to each other in appearance. Song similarity as a standard of species identity also applies to nesting cuckoos.African Black Coucal Centropus grillii have been considered conspecific with Lesser Coucal C. bengalensis of Asia and Madagascar Coucal C. toulou of Madagascar, but their voices differ as do their size and plumage, so they are now regarded as distinct species. The C. grillii song is “kop-kop” repeated at 2-sec intervals, sometimes in a long series, each note rising and

Figure 6.5. Brush Cuckoos Cacomantis variolosus: Northern populations, a–b, “where’s the tea” (a, Gombak, Malay Peninsula, NSA 32539; b,Tegah, Sulawesi, NSA 32750); c–d,“fear-fear” (c, Lore Lindu NP, Sulawesi, NSA 13137; d, Mt Katanglad, Mindanao, NSA 41264); southern populations, e–g,“where’s the tea” (e, f, FitzRoy River,Western Australia, Buckingham and Jackson 1990; g, Mt. Hagen, New Guinea, NSA 38418); h–i, “fear-fear” (h, Morobe, New Guinea, NSA29701; i, Lamington NP, Queensland, NSA14259). Note: time scale 0–4s correct for a, b, e–g; time scale to read 0.8s for c, d, h, i.

Species 103

Figure 6.6. Little Bronze-cuckoo Chrysococcyx minutillus: a–g, descending trill (a, jungei, Sulawesi, Ben King; b, crassirostris, Tanimbar, NSA 46764; c, poecilurus, Timor, NSA 72300; d, poecilurus, Timor, Ben King; e, minutillus, Kimberley Range, WA, Graeme Chapman; f, poecilurus, Cardwell, QD, Buckingham and Jackson 1990; g, barnardi, Brisbane, QD, Barry Morgan); h–n, whistled song (h, peninsularis, Malay Peninsula, Scharringa 1999; i, Sulawesi, Ben King; j, Flores, NSA 72301; k, Kimberley Range,WA, Graeme Chapman; l, Darwin, NT, Buckingham and Jackson 1990; m, Cardwell, QD, Buckingham and Jackson 1990; n, Brisbane, QD, Barry Morgan).

falling in pitch within the range 0.8–1.0kHz, and the notes are repeated 6 per sec for a couple of seconds with no change in pitch. C. toulou gives a muffled series of 5–10 hoots “toulou toulou toulou . . .” decreasing in volume, short notes given c. 6 per sec at 0.6kHz, and it also gives water-bottle call duets, one bird in a “toulou” series, the mate joining with a higher-pitched series at 14 notes per sec dropping in pitch and slowing to 8–10 notes per sec; while the Asian Lesser Coucal has paired notes lower in pitch, and not mellow and modulated as in the African Black Coucal. In Figure 5.3, these three coucals are not each others’ closest relatives, insofar as the species complex includes the Philippine Coucal C. viridis which is sympatric with C. bengalensis in the Philippines. The differences in song are consistent

with morphology and results of molecular genetics which show that these coucals are distinct species. In other cases, sympatric, non-interbreeding coucal species may sound alike.The pitch of the call differs between the sexes; the larger female has a lower-pitched call, noticed in duetting pairs. Taking into account the pitch differences between sexes and the variation in calls given by a single bird, the geographically remote populations of Greater Coucal Centropus sinensis in southern Asia and the Andaman Islands appear to be the same species, although these birds differ in size and sometimes in plumage.The sex of the bird was not noted in field recordings, and measurements of the pitch of calls in the coucal species accounts may not be comparable where a male was recorded for one species and a female for

104 The Cuckoos another species. Chappuis (2000) described interspecific song duets between Senegal Coucal Centropus senegalensis and Blue-headed Coucal C. monachus, and song duets between Black-throated Coucal C. leucogaster and C. monachus, and Brosset and Erard (1996) described duets between Gabon Coucal C. anselli and C. monachus. Also the calls of C. leucogaster in Nigeria and C. anselli in Congo are similar to each other and to calls of C. monachus, and coucals sometimes react to the playback of recorded calls of other coucal species much as they do to calls of their own species (Dowsett and Dowsett-Lemaire 1993). Song of yellowbill malkohas Ceuthmochares vary together with the plumages of malkohas across Africa and this covariation indicates there are two species,

rather than one species. Across West Africa from Guinea-Bissau to Ivory Coast, Chattering Yellowbill C. aereus have a monotonic chatter that quickens to a rattle; from Nigeria and Cameroon eastward to western Kenya, they have a similar quickening rattle but the song often begins with a high pitch an octave above the underlying basic pitch and it drops more than an octave to the terminal rattle. In contrast,Whistling Yellowbill C. australis from coastal Kenya to South Africa have a song with a set of long whistles interposed between the introductory notes and the terminal rattle (Figure 6.7). In southeast Asia, Chestnut-breasted Malkohas Phaenicophaeus curvirostris present a problem of species distinctiveness. The plumage pattern, bill

Figure 6.7. Ceuthmochares African yellowbills. a-f, Chattering Yellowbill C. aereus (a, Bijagos Archipelago, GuineaBissau, C. R. Barlow; b, N’Douci, Ivory Coast, Chappuis 2000; c, Mt Bengoué, Gabon, Chappius 1974; d, Obale, west of Lake Kivu, NSA 80609; e, Nyungwe Forest, Rwanda, NSA 30638 cf. Dowsett-Lemaire 1990; f, Kakamega forest,W Kenya, NSA 48766); g-i, Whistling Yellowbill C. australis (g, Sokoke forest, E. Kenya, C. R. Barlow; h, Mrima hill, SE Kenya, NSA 18022; i, Kruger NP South Africa, Gillard 1987).

Species 105 pattern, and the shape of the nostril and groove on the bill of these malkohas differ among islands across the western Sumatran islands and Greater Sunda Islands to Palawan. It is unknown whether songs differ between these malkoha populations. Questions about the species of cuckoos remain to be investigated. Fieldwork in India and Southeast Asia is needed to test whether both kinds of song occur in extensive sympatry in the drongo-cuckoos Surniculus dicruroides and S. lugubris. In Australasia, questions of song and species distinctiveness need to be studied in detail in the Koels Eudynamys scolopacea, in the hawkcuckoos Hierococcyx sparverioides, H. bocki and H. varius, in Brush Cuckoos Cacomantis variolosus and in Little Bronze-cuckoos Chrysococcyx minutillus. Finally, additional field studies are needed on several Asian Cuculus cuckoos.The songs of three species, Oriental Cuckoo Cuculus optatus, Himalayan Cuckoo C. saturatus and Sunda Lesser Cuckoo C. lepidus differ from each other; their plumages differ only in size and color of the underparts (and juvenile of C. optatus differ in plumage from the other two). Here they are recognized as three distinct species, as in King (2005).

Speciation New species originate when their common ancestral populations are isolated over time and evolve independently, with genetic differences accumulating to the degree that the populations become reproductively independent (Mayr 1942, 1963). Geological changes that divide birds into geographically isolated populations (vicariants), and dispersal of individuals from a region into new areas where the birds breed separately from their source populations, are two processes that can isolate these populations (Zink 1996, Klicka and Zink 1997, Zink et al. 2000). On the other hand, speciation in the brood-parasitic cuckoos might involve another process, where speciation of the cuckoos is closely linked to the speciation of their hosts. This section develops the idea that allopatric speciation accounts for the differentiation of both nesting cuckoo species and of the brood-parasitic cuckoos.

Geographic speciation model The idea of recent speciation occurring in vicariant populations that live in isolated areas leads to a pre-

diction of geographic complementarity, much as in Mayr (1963).The extent to which regional geological changes have led to speciation is tested by finding common biogeographic patterns across different lineages (Cracraft 1985, 1988).Vicariance biogeography is based on a congruence of historical branching sequences among clades. The reasoning is that regions became geographically disjunct at the time of unique geological events that shaped the face of the earth’s landscape, and the isolated populations diverged into species (Haffer 1978, 1992, 1997, Cracraft 1982, Cracraft and Prum 1988, Marks et al. 2002). The effect of dispersal on geographic isolation is clear on islands that were never connected to continental areas, and closely related species on islands and mainland were never in contact. These birds must have dispersed from mainland to island. In many cuckoos the geographical distribution of closely related species is complementary, where the species breed only in neighboring regions where the other does not occur. Cuckoo species with distributions consistent with this model include brood-parasitic hawk-cuckoos Hierococcyx hyperythrus, H. nisicolor, H. fugax and H. pectoralis in Asia, the drongo-cuckoos Surniculus in southern Asia, the plaintive cuckoos Cacomantis merulinus and C. passerinus in southern Asia, and the long-tailed cuckoos Cercococcyx olivinus and C. montanus in Africa. Other sets of closely-related species that live in geographically separate areas include the nesting cuckoos such as lizard-cuckoos Coccyzus (“Saurothera” and “Hyetornis” ) in the West Indies, the ground-cuckoos Carpococcyx in southern Asia, ground-cuckoos Neomorphus in the New World tropics, and coucals Centropus in the New Guinea region. In South America, the ground-cuckoos Neomorphus live in regions that are now separated from each other by large rivers. Haffer (1977) proposed that ground-cuckoos vary in morphology in the way predicted by a series of climate changes with dry periods that separated the formerly continuous continental forests, with the forest patches persisting in refugia that became separated by large expanses of savanna habitat. First a speciation event separated the ancestors of radiolosus and geoffroyi from the ancestors of pucheranii and rufipennis, then each of these groups speciated once more. Ground-cuckoos could not cross from one forest refuge to another over broad

106 The Cuckoos rivers or across broad areas of savanna. The phylogeny of Neomorphus is consistent with this model. N. geoffroyi has subspecies both north and south of the upper Amazon in Brazil; in the lower Amazon ground-cuckoos are not known north of the river in Pará and Amapá; they occur south of the river. N. pucheranii occurs on both banks of the upper Amazon, with different subspecies on each bank as far towards the headwaters as they have been sampled. N. radiolosus is separated from other groundcuckoos by the Andes, a distribution that suggests the cuckoo was isolated from N. geoffroyi with the rise of the northern Andes. N. rufipennis occurs in the northern drainage basin east of the Orinoco and west of the Courantine rivers, and its restricted distribution suggests the Courantine is a barrier to dispersal into Suriname; the species also lives on the upper Rio Branco in the northern Amazon basin. Dispersal in a continental area is likely to lead to gene flow and a breakdown of local differentiation. Wallace (1853) observed that certain mammal species replace one another on opposite banks of the large rivers in Amazonia. A river of a kilometer or more in width may prevent dispersal and genetic interchange of forest birds with limited abilities of flight.At the same time, the continuous forests along a river may allow dispersal and gene flow upstream where the river narrows in the headwaters and is no longer a barrier to dispersal, and the birds can disperse over or around the river and downstream again on the opposite side. Wallace noted, “Toward their sources, rivers do not form a boundry between distinct species.” In time, populations may become separated when the river changes course, and populations on each side of the river become isolated. The idea is that populations differentiate into species when barriers arise at the periphery of their range, but these barriers are effective only where they prevent dispersal, and the isolation of populations is not caused by dispersal. Dispersal would lead to speciation only where occasional movements of individuals leads to the establishment of new geographic isolates across regions that more often bar dispersal. Changes in climate during the Pleistocene, the rise of the Andes, and changes in river courses in the western Amazon with a breakdown of distinction between flooded várzea forest and non-flooded tierre firme forest repeated over time by means of channel

migration and sedimentation of the meandering rivers (Räsänen et al. 1987) may all have affected these Amazonian birds.The changes might isolate populations at one time, and at other times they might allow them to disperse, interbreed and exchange genes again. In Neomorphus there is biogeographic support for a vicariance model of speciation. Nevertheless it is unclear whether the geographic isolation of populations in separate regions of relict forests when these were separated during a Pleistocene dry period was involved in the initial separation and speciation, or whether the forest remained intact over a long period and the initial isolation of the populations occurred when a new broad river divided a formerly continuous forest. Because more than one geological process can account for the present-day distribution of the ground-cuckoo species, neither process is excluded, and the ecogeographical processes that were responsible for the speciation of these ground-cuckoos remain uncertain. Recent history witnesses rivers as barriers to dispersal in the short term, if not in speciation. Ground-cuckoos on Barro Colorado Island disappeared a few decades after the island was isolated from mainland populations. Birds fly, and they disperse, and as a result, dispersal is important in the distribution of many species. Many cuckoos are strong fliers and move thousands of kilometers each year. In the New World, the closely related Yellow-billed Cuckoo, Pearl-breasted Cuckoo and especially the insular Mangrove Cuckoo Coccyzus species likely trace their geographic distribution in the breeding season to dispersals that are independent of major geological change, insofar as all three cuckoos are seasonally migratory. In the Old World, the hawk-cuckoos Hierococcyx hyperythrus and H. nisicolor migrate from their northern breeding areas into tropical regions where other hawk-cuckoos H. fugax live year-round. And yearround resident malkoha species live on small islands well offshore as well as on the mainland of Asia and the Greater Sunda Islands. The coucals Centropus are of interest for models of allopatric speciation. These birds live in a region of active geological history where both geological processes of landscape change and the biological process of dispersal have been involved in isolation and speciation. Movements of the Australian and

Species 107 Philippine-Caroline continental plates present a complex background to a vicariance model of speciation. In the Tertiary, Australia and southern New Guinea were on a shallow sea plate along with the Aru Islands. Halmahera and the other islands of the northern Moluccas were not part of this block. The Banda Sea arc including the Kai Islands as its eastern bend had a separate origin from the New Guinea plate; these western islands were never connected with Australia and New Guinea (Hall 1996, 1998, Milson et al. 1996). In the genetic phylogeny, Goliath Coucal Centropus goliath of the northern Moluccas is not closely related to Greater Black Coucal C. menbeki of New Guinea, as proposed by White and Bruce (1986). Rather, C. goliath is most closely related to Pheasant Coucal C. phasianinus, which occurs in Australia, New Guinea, the Kai Islands (C. p. spilopterus) and Timor (C. p. mui). The geological history of land masses suggests that dispersal as well as a vicariance isolation of populations by land movement led to speciation in these coucals. The same argument applies to the question of whether C. goliath is closely related to either species. In both cases a dispersal would be required to get the birds from New Guinea to the northern Moluccas. We expect that episodic dispersal has led to the establishment of populations in new areas, where they diverged into species that differed from their source population. Even short-winged coucals disperse, as one (African Black Coucal Centropus grillii ) is a seasonal migrant.There is historical evidence of success in dispersal on a broad geographic scale. Krakatau Island, between Sumatra and Java, blew up in a devastating volcanic explosion in 1883 and all land birds and land plants were exterminated (Dammerman 1992, Thornton 1996). Lesser Coucals C. bengalensis appeared there by 1908 and nested soon afterwards; Greater Coucals C. sinensis were common by 1919. Greater Coucals occur on offshore islands in the Indian Ocean, and they have historically spread into northern India. Common Koel Eudynamys scolopacea have colonized offshore islands near New Guinea after the islands lost all their animals in volcanic explosions about 300 years ago. In India the resident Sirkeer Malkoha Taccocua leschenaultii spread through dry scrubland as they dispersed along irrigation canals in the past century. In the New World, Groove-billed Ani Crotophaga sulcirostris established themselves in

Florida after a hurricane in 1926, while other populations have come, gone and come again in dispersal with loss in storms and recolonization from other islands in the West Indies, and in Central America the anis enlarged their distributional range when forests were cleared and the birds dispersed from one open habitat to another.

Host-parasite cospeciation model The close association between a brood parasite and its host species might provide an opportunity for reciprocal evolution, or coevolution. The cuckoo evolves mimicry of its own egg to the host egg, and the host might evolve to recognize a cuckoo egg and reject it from the nest. Or the parasite-host association might lead to speciation of cuckoo populations matched with the speciation of their host populations, in parallel speciation or cospeciation. The brood-parasitic cuckoos as a group are not hostspecific, as far as we know. Only a few cuckoo species are restricted to a single kind of host. In many cuckoos the hosts are not well known, and further fieldwork may reveal a broader range of their host species. Perhaps the most host-specific cuckoo is Little Bronze-cuckoo Chrysococcyx minutillus which uses flyeaters of the genus Gerygone. These cuckoos have not undergone speciation in parallel with their hosts. The problem of host specificity and speciation in these birds involves determining not only the songs and genetic histories of cuckoos but also the species limits in Gerygone. Sulphur-bellied Flyeater G. sulphurea in the Malay Peninsula, Java and Flores comprise a single species with no morphologically distinct geographic populations or subspecies, whereas the cuckoos differ geographically in these regions. Among the cuckoos, the bronze C. m. poecilurus parasitizes Large-billed Flyeater G. magnirostris in New Guinea and northern Queensland, and the greenish C. m. minutillus in northeastern Queensland are in more open habitats where White-throated Flyeater G. olivacea may be the host species.The same cuckoos also parasitize sunbirds in New Guinea and Australia. Also, Shining Bronze-cuckoo parasitize two species of Gerygone in New Zealand and the Chatham Islands. These populations show a lack of cospeciation in the cuckoos and their songbird hosts.

108 The Cuckoos In a second case,Thick-billed Cuckoos Pachycoccyx audeberti in Africa use host helmet-shrikes Prionops, although they must use other hosts in Madagascar where helmet-shrikes do not occur.The host species are more closely related to other songbirds (insofar as we can compare genetic distances across different genes, Cibois et al. 1999, 2001,Yamagishi et al. 2001, Sefc et al. 2002), than Thick-billed Cuckoos are to other cuckoos. Furthermore, the subspecies of Thick-billed Cuckoos within Africa do not coincide geographically with species of their helmet-shrike hosts (Fry et al. 2000). In a third case, African Cuckoos Cuculus gularis in most parts of Africa use Fork-tailed Drongos Dicrurus adsimilis although in west Africa they also use Yellow-billed Shrikes Corvinella corvina, which being local as hosts may be secondarily-acquired hosts. It appears that brood-parasitic cuckoos have switched from time to time from old hosts to a new host species. In these cuckoos, the alternate host species are not closely related to the primary host species. First, in the Indopacific region, Common Koels have switched hosts from crows Corvus spp. to parasitize mynahs Acridotheres spp. in southern Thailand and in the Malay Peninsula within the past century, as one host became scarce and the other became more numerous with changes caused by human alteration of the habitat.The variation within time suggests a number of host switches in the past in this species, which parasitizes mynahs elsewhere in Indonesia and large honeyeaters in Australia. Great Spotted Cuckoos have begun to parasitize certain host species in recent years.The cuckoos now parasitize House Crows Corvus splendens in Israel, with records beginning in 1997 (Yosef 1997, 2002). The cuckoo and the host have only recently come into contact. House Crows spread from India to Egypt and the east coast of Africa in the first half of the twentieth century, and they have bred in the Middle East only since the 1970s (Shirihai 1996).The cuckoos and House Crow were not in sympatry in earlier years and had no earlier opportunity to interact.These cuckoos parasitize other species of crows in the Middle East, Egypt and around the Mediterranean, so colonizing the new host did not involve a great change in cuckoo behavior. Also, Great Spotted Cuckoos parasitize Common Mynah

Acridotheres tristis near Estcourt and Colenso, Natal, South Africa (Symons 1962, Sharp 1976). Mynahs in South Africa were introduced from Asia in 1888 to 1900 (Maclean 1993). The cuckoos in Weenen County regularly parasitize open-nesting crows and hole-nesting starlings (West et al. 1964), and mynahs nest in the same kind of holes in buildings and trees as the native starlings. Certain Asian cuckoos parasitize two or more closely related species of hosts even though these cuckoos are not all each others’ closest relatives. In this case the five cuckoos Cuculus poliocephalus, C. canorus, C. optatus, C. lepidus and C. saturatus each parasitize small warblers of the genus Phylloscopus, and these cuckoos sometimes parasitize the same host species. These cuckoo-host associations indicate that these hosts have been colonized more than once by more than one kind of cuckoo (Becking 1981, Higuchi and Sato 1984). The same cuckoo species also parasitize other species of hosts that are not closely related to Phylloscopus warblers.These associations of brood-parasitic cuckoos and hosts tell a history of independent colonization by cuckoo species of their host species, rather than speciation of cuckoos that went along with the speciation of their hosts. In all these associations, the brood-parasitic cuckoos and their hosts have speciated independently. There is no evidence that cuckoos and their hosts have cospeciated, or that phylogenetic relationships among the hosts give clues to the relationships among the cuckoos. In their independent histories of speciation, the parasitic cuckoos are similar to the African finches. Among the finches, the brood parasites and their hosts speciated along completely independent lines where one tree of life did not shadow the other. Genetic differences between closely related viduid species are less than the genetic distances between the corresponding host species even though both lineages have the same generation times. The genetic phylogenies point towards an evolutionary history of successful switches of a line of parasites from an old host to a new host (Klein and Payne 1998, Sorenson and Payne 2001, 2002, Payne et al. 2002). In summary, geographic isolation has led to speciation in cuckoos, much as it has in other birds (Mayr 1963). Still, it is curious that the cuckoo species which are genetically most similar to each other are brood parasitic, particularly the Cuculus cuckoos.

7 Fossil and comparative evidence of cuckoo relationships Fossil cuckoos In the early Tertiary (Houde and Olson 1988, Mayr 1998a,b), the oldest fossil birds that might be cuckoos are known in the Paleocene and Eocene, the time when many avian families first appeared in the fossil record. None of these ancient fossils are known for certain to be cuckoos (Feduccia 1996, G. Mayr, in litt.). An Eocene fossil Foro panarium Olson 1992 from Wyoming is a nearly complete specimen that has some cuckoo-like characters such as a pectineal process of the synsacrum. Foro is said to represent a primitive group of land birds (Olson 1992). Cuckoos are not unique in having a pectineal process, as it also occurs in other terrestrial birds, and Foro may not be a cuckoo. Fossils from the early Tertiary have been identified as cuckoos from a tarsometatarsus or humerus. Marsh (1872) described a small fossil, Uintornis lucaris, from the Eocene of Wyoming as a woodpecker, on the basis of the distal end of a tarsometatarsus. The fossil was illustrated and described in detail by Shufeldt (1915), who thought it was not zygodactyl and it was neither woodpecker nor cuckoo. Brodkorb (1971) listed it as a cuckoo; others listed it in a fossil family Primobucconidae (Feduccia and Martin 1976, Olson 1985). Fossil tarsometatarsi from the London Clay deposits were the material used to describe two other small birds, Parvicuculus minor and Procuculus minutus (Harrison and Walker 1977). Olson and Feduccia (1979) noted that cuckoos have a posteromedial twist of the outer trochlea (number IV), and this twist was absent in the

London Clay specimens. Harrison (1982) attributed the lack of a distinctive trochlea IV to a “generalized” cuculiform condition and cited the lack of the feature in turacos and hoatzin as evidence, but the assumption that turacos and hoatzin are cuculiform birds is questionable. Harrison (1982) described several features of “cuculiform” tarsi, and proposed a new family “Parvicuculidae” to accommodate the London Clay fossils with the view that they were not cuckoos in the modern sense. Baird and Vickers-Rich (1997) described a fossil tarsometatarsus from the Late Paleocene of Brazil as a cuckoo, Eutreptodactylus itaboraiensis. In this fossil, trochlea IV is large, rounded in shape, lacks a sulcus and lacks a distinct sehnenhalter, and trochlea IV is rotated only slightly from the axis formed by the distal projections of trochleae II and III, whereas it is rotated more than 60° in modern cuckoos and other zygodactyl birds, so it is questionable that the fossil tarsometatarsus is of a zygodactyl bird. Finally, Chandler (1999) described Eocuculus cherpinae from Colorado (Chandler 1999), complete except for a skull. The distal part of the tarsometatarsus may have a large trochlea IV with an accessory articulating process or sehnenhalter. This process is like that of cuckoos insofar as the process in cuckoos is weakly developed, yet the bone of Eocuculus is not clearly zygodactyl, and the similarity with cuckoos is the lack of a diagnostic sehnenhalter such as occurs in the other modern groups of zygodactyl birds. In these Tertiary fossils the identification of a tarsometatarsus as a cuckoo was “cuculiform by elimination, and more similar to Cuculidae than Musophagidae or Zygodactylidae” (Harrison and

110 The Cuckoos Walker 1977). One diagnostic definition of a cuckoo metatarsus is “elongate, narrow corpus tarsometatarsus, lenticular-shaped trochlea metatarsi IV lacking any sulcus” (Baird and Vickers-Rich 1997). But some cuckoos (Scythrops) have a broad tarsometatarsus. The other half of the definition, like the descriptions of Harrison (1982), reflects the lack of distinctive features of a cuculid tarsometatarsus. The other skeletal element used to identify early Tertiary fossils as cuckoos is the humerus. Dynamopterus velox Milne-Edwards 1892 in the EoOligocene chalks of southern France was described from a humerus but with no comparison with modern cuckoos other than in size. The next candidate for oldest cuckoo is Neococcyx mccorquodalei Weigel 1963 from the Lower Oligocene Cypress Hills in Saskatchewan, from a humerus that differed from modern cuckoos in having the entepicondyle reduced, in having the ectepicondyle ridge less angular, and in other features of shape. The humerus was about the size of the humerus of Coccyzus; the internal condyle was smaller and the olecranal fossa was more shallow (Weigel 1963, Martin and Mengel 1984). Because these authors did not describe differences in the humerus of cuckoos and other birds, the identification of fossil Dynamopterus and Neococcyx as cuckoos is not well supported. Fossil cuckoos are known from the middle to late Tertiary and the Quaternary in the Northern Hemisphere, and in the late Quaternary in the Southern Hemisphere. The fossils establish that cuckoos were present and were similar in morphology to modern cuckoos. A fossil cuckoo Cursoricoccyx geraldinae Martin and Mengel 1984 from the early Miocene c. 20 mya (million years ago) of Colorado was smaller than the modern Lesser Roadrunner Geococcyx velox, and had a less strongly developed coracoid, and tarsometatarsal elements in specialization for cursorial life; the carpometacarpus was longer than in modern terrestrial cuckoos. Fossil roadrunners, perhaps the same species as Greater Roadrunner G. californianus, and a larger one known as Conkling’s Roadrunner G. conklingi (the two may be conspecific), are known from the Pleistocene. They per-

sisted into the Holocene to less than 6000 years ago in arid regions of the southwestern United States and Mexico (Larson 1930, Howard 1931, Miller 1943). G. conklingi was larger and may have been adapted to inland temperatures and a continental climate (Harris and Crews 1983). In Australia a fossil coucal Centropus colossus, known from a humerus, was larger than the presentday Australian Pheasant Coucal C. phasianinus. The fossil coucal lived within the past 40,000 years (Baird 1985a,b, Rich and Baird 1986). In Madagascar, a large fossil coua Coua primavea is known from the Quaternary, and a larger subfossil coua C. berthae is known from recent cave surface deposits (Milne-Edwards and Grandidier 1895, Goodman and Ravoavy 1993, Burney et al. 1997), along with several extinct or locally extinct birds and mammals. C. primavea had a gracile tarsometatarsus of 84 mm, longer and more slender than that of C. delalandei and C. gigas; C. berthae had a broad pelvis and a robust tarsometatarsus of 93 mm, longer and more robust than any other coua. C. berthae may have been flightless.These fossil couas lived 1000 years ago and perhaps as recently as 600 years ago. Remains of other cuckoos in the upper layers of the deposits include the extant Madagascar Lesser Cuckoo Cuculus rochii, Madagascar Coucal Centropus toulou, and other couas including Giant Coua Coua gigas.The associated human artifacts appear to be more recent than the extinct fossil couas (Burney et al. 1997). Finally, on St. Helena Island a small bird Nannococcyx psix was described from an unmineralized partial humerus, identified as a cuckoo and dated as later than 1502, when the island in the South Atlantic Ocean began its ecological degradation with the impact of man and the loss of its forests (Olson 1975, 1990, Rowlands et al. 1998). Recent and prehistoric remains of extant cuckoo species in the Old World include Great Spotted Cuckoo in Israel and Common Cuckoo in Europe, and in the New World the Yellow-billed, Mangrove and Black-billed Cuckoos Coccyzus, two squirrel cuckoos Piaya, and a lizard-cuckoo Coccyzus (“Saurothera”) in the West Indies, two anis Crotophaga in the West Indies and Yucatan, American Striped Cuckoo Tapera naevia in Brazil,

Fossil and comparative evidence of cuckoo relationships 111 and roadrunners in the North American southwest (Brodkorb 1971, Olson and Hilgartner 1982, Morgan 1994). Sites in Melanesia include remains about 8000 years old of Brush Cuckoo Cacomantis variolosus and coucals Centropus thought to be extant species (Steadman et al. 1999); and Longtailed Cuckoo Urodynamis taitensis is known as a subfossil on Henderson I in the Pitcairn Islands from prehistoric Polynesian occupation c. 400– 1000 years ago (Weisler 1995, Wragg 1995), and from Tonga c. 400–2800 years ago (Steadman et al. 2002). In summary, the recent fossil cuckoos show a greater diversity of couas than now live in Madagascar.The fossils do not suggest a great loss of birds caused by human immigrations such as happened in island avifaunas of the Pacific (Olson and James 1982, 1991, James and Olson 1991, Collar et al. 1994, Steadman 1995). The fossils also show extant cuckoos in the same areas where they now occur in the Old and New Worlds.

Cuckoos and the clocks of evolutionary time Estimates based on molecular genetics, of the time of separation of the major groups of birds tend to be older than estimates from the fossil record (Sibley and Ahlquist 1990, Feduccia 1996, García-Moreno and Mindell 2000, van Tuinen and Hedges 2001). The former suggest divergence of modern orders of birds in the Cretaceous, whereas the earliest fossils of the same orders are known from the Tertiary. The first estimate of an early separation of the major groups of cuckoos was that of Sibley and Ahlquist (1990) at 69 million years ago.Their estimate was based on their values for molecular thermal dissociation of Old World and New World cuckoos, compared with the value of thermal dissociation between Ostrich Struthio camelus and Greater Rhea Rhea americana, and a time calibrated from the separation of continental plates by the Atlantic Ocean. Another molecular estimate of the time of separation of Old World and New World cuckoos is 52 million years ago, with a calibration based on a fossil anseriform Presbyornis and a divergence of Ostrich from other ratites at 80 million

years (van Tuinen and Hedges 2001). Other estimates of the time of separation of Africa and South America are 100–120 million years (Hay et al. 1999), somewhat earlier than the estimate used by Sibley and Ahlquist. A more recent possible touchstone is the collision of tectonic plates of the Bismarck Archipelago and New Guinea about 10–15 million years ago (Lee and Lawver 1995, Michaux 1998, Polhemus and Polhemus 1998); coucals in these areas are each others’ closest relatives and may have diverged from a common ancestor sometime after this tectonic event. The basal phylogenetic split of Old World and New World cuckoos (Figure 5.1) suggests an evolutionary history of cuckoos dating to before the separation of Gondwana, or sometime earlier than 60 million years ago, as in other Gondwana clades (Cracraft 2001). Genetic differences between Old World and New World cuckoos also suggest a divergence around 60 million years ago (van Tuinen and Hedges 2001). The estimate of age of 53 million years for the oldest fossil cuckoo (Feduccia 1996) was based on an undescribed bone of a bird of uncertain relationships. Thus, biogeography and molecular estimates point to an origin of cuckoos somewhat earlier than the known fossil record.The genetic and the fossil estimates agree that cuckoos have been a distinct lineage for tens of millions of years. The most secure conclusion is that cuckoos were around after the Cretaceous and diverged during the Tertiary.

What birds are related to the cuckoos? From a historical perspective, the ideas about the relationships of cuckoos are seen in the names given when cuckoos were originally described as new species and were thought to be members of other families of birds. Channel-billed Cuckoo Scythrops novaehollandiae was once known as “Anomalous Hornbill” and “Psittaceous Hornbill”,Asian Emerald Cuckoo Chrysococcyx maculatus was described as a trogon Trogon maculatus, and Pallid Cuckoo was described as a dove Columba pallida. Conversely, Greater Honeyguide Indicator indicator was described by Sparrman as Cuculus indicator, and Green Turaco

112 The Cuckoos Tauraco persa was described by Linnaeus as Cuculus persa. In addition, several birds of other families were named at the species level as cuckoo-like: an owl, Asian Barred Owlet Noctua (Glaucidium) cuculoides and a hawk, African Baza or Cuckoo-falcon Aviceda cuculoides (Sibley and Monroe 1990). Common names reflect the similarity in appearance of cuckoos to other groups, as Cuckoo-doves Macropygia, Cuckoo Rollers Leptosomas discolor and cuckooshrikes Campephagidae in the Passeriformes. Other cultures likewise have remarked on the resemblance of cuckoos to owls and hawks, as in the common names of these birds in the Moluccas (Taylor 1990). Traditional systematic views of avian relationships are based on morphology and suggest that ratites and tinamous are basal to all other birds (Huxley 1867, 1868, Fürbinger 1888). Recent molecular systematics studies support (a) this conclusion, (b) that waterfowl and Galliformes are each other’s closest relatives, and (c) that these are basal to the remaining groups of birds (Groth and Barrowclough 1999, van Tuinen et al. 2000, Johnson 2001, Sorenson et al. 2003b), but they do not resolve the relationship of the cuckoos with these other birds. One recurring idea in avian systematics is that cuckoos are related to the Neotropical Hoatzin Opisthocomus hoazin (Walters 2003). In their anatomical work, Huxley ( 1867, 1868) and Fürbinger ( 1888) found hoatzin most similar to the Galliformes. Sibley and Ahlquist ( 1972) cited Huxley as listing hoatzin “adjacent to the Coccygomorphae”; in fact Huxley concluded that hoatzin is related to galliform birds and doves, and he did not list the cuckoos and hoatzin next to each other. In an anatomical survey of avian leg muscles, the cuckoos and hoatzin were said to be each others’ closest relatives (McKitrick 1991), but the survey included only those cuckoos with formula leg muscles like those of hoatzin and it excluded cuckoos with other leg muscles (e.g. Berger 1952, 1960). In a comparison of egg-white protein electromorphs, hoatzin was more like crotophagine cuckoos than were other cuckoos, and Sibley and Ahlquist (1972, 1973) concluded that hoatzin was a crotophagine cuckoo (however, the results could not be repeated: Brush 1979); and in pairwise tests of molecular thermal association (DNA-DNA

“hybridization”), hoatzin was unlike the cuckoos, and except for Galliformes no other birds were included in the comparison (Sibley and Ahlquist 1990: Fig. 79). More recent, sequence data show that cuckoos and hoatzin are not related (Sorenson et al. 2003b). Cuckoos share several morphological features with turacos, and the cuckoos and turacos are sometimes listed together in a large single order Cuculiformes. This convention was most strongly established by Fürbinger (1888), the first avian anatomist to present a comprehensive phylogenetic hypothesis based on discrete characters. Cuckoos (Sorenson et al. 2003b) and turacos shared more morphological characters (“Merkmale”) with each other than either family did with any other group of birds. Fürbinger’s Tables xli and xlii listed 46 morphological characters (“Merkmale”) including the skeleton, muscles and feather tracts in the families and orders of birds, and features of birds, based on his own anatomical studies. A total of 32 out of these 46 characters were shared by cuckoos and turacos; though a few characters were inconsistent within one family, many characters were also shared with many other avian taxa, and none of the 32 characters were shared exclusively between cuckoos and turacos.The most nearly similar traits shared nearly exclusively by the two groups were not discrete characters, but were the relative lengths of the wing bones and the leg bones (Fürbinger’s characters 20, 21 and 23). Cuckoos are also similar in body form and plumage to doves Columbidae, and their leg muscles and tendons are like those of some parrots Psittacidae (Raikow 1985). In some features of the skull, Hoatzin is similar to cuckoos, in others to turacos, and it is not similar to the curassows Cracidae (Marceliano 1996).The palate of hoatzin is schizognathous, like the palate of Galliformes and some Gruiformes such as the Neotropical trumpeters Psophiidae, and unlike the desmognathous palate of turacos and cuckoos (Huxley 1867, Sibley 1955, Sibley and Ahlquist 1990, LSU, UMMZ). Another idea is that cuckoos are related to both turacos and hoatzin. One molecular study found hoatzin more closely related to turacos than to cuckoos, and these three groups formed a clade

Fossil and comparative evidence of cuckoo relationships 113 (Hughes and Baker 1999). However, only a few other avian orders were used in that comparison, and more extensive comparisions failed to support those two conclusions (Sorenson et al., 2003b). Finally, a preliminary comprehensive analysis of skeletal characters located the hoatzin, turacos, and cuckoos each at different regions of the tree of avian relationships, with cuckoos most closely related to certain coraciiform birds, hoatzin to gruiforms, and turacos to a wide assemblage of orders

including mousebirds Coliiformes, rails Gruiformes, woodpeckers Piciformes and passerines Passeriformes (Livezey and Zusi 2001). Sibley and Ahlquist (1990) concluded that “DNA comparisons show that the cuckoos have no close living relatives.” Mitochondrial and nuclear gene sequences also indicate no close relatives (Sorenson et al. 2003). These genetic studies suggest that we know no single order of birds that is most closely related to cuckoos.

8 Breeding biology and life histories

The breeding behavior of cuckoos ranges from solitary pairs, to social groups that breed cooperatively, to brood parasitism. Cuckoos are altricial, and the young hatch and remain in the nest in a nearly helpless state, unable to feed themselves, and are completely dependent on their parents or their foster parents to provide care and to feed them. The adult females vary from providing parental care to their young in their nest, to the absence of both nest-building behavior and parental care in the brood-parasitic cuckoos. The adult males take on the life style of their females, and some males even provide more parental care than their mates. Among nesting cuckoos, many are secretive and solitary and live alone or in breeding pairs. In contrast, the communally nesting anis and Guira Cuckoos are conspicuous and social birds, living in the same social group both in the breeding season and also when they are not breeding.

Breeding displays Breeding displays are known for only a few cuckoos. A few common themes are seen when the male displays in courting the female. The tail is raised towards the female, and the male offers a bit of food to the female he courts.The display behavior of the Greater Roadrunner of North America, a nesting species, and the Common Cuckoo of the Palearctic, a brood-parasitic species, both of which have been closely observed in the field can be taken to represent that of cuckoos in general. In Greater Roadrunners the male and female spend time foraging together and calling back and forth before they show a sexual interest in each

other. Early in courtship, one bird runs after the other on the ground, often for several hours; both birds stop and rest between chases.The chasing bird runs and lunges at the forward bird with its wings and tail raised and fanned, while both birds give “clack” calls. The male gives a “coo” call from an elevated perch. One bird carries a stick in its bill and presents it to the mate; both sexes pick up sticks and pass them to the mate or drop the stick at the mate’s feet. The behavior may provide the mate with an assessment of its partner before it commits to a nest. Before they copulate the roadrunners exhibit with a “prance display.” The male approaches the female on the ground in a short burst of speed, holds an insect or snake or lizard in his bill, and wags his tail back and forth. The male runs to and from the mate with its wings and tail lifted, then he lowers his wings and brings them inward with a “pop.” He holds his tail over his body, then gradually lowers his tail, exposes the postorbital bare skin, erects his crest, and sleeks the contour feathers. The display involves four or five cycles of lifting the wings: it lasts more than two minutes. In a “tail-wag display” the male wags his tail from side to side while he bows, then slowly lifts his head, as he faces the female and gives a “whirr” call. After this he jumps into the air and leaps over the female or mounts her from the rear. He holds the food in his bill and presents it to the female as they copulate. Then he dismounts, they walk away from each other, they flick the tail, and the female eats the offering or feeds it to her young if copulation occurs after the young have hatched: roadrunners copulate socially, and not only for fertilization (Whitson 1975).

Breeding biology and life histories 115 The displays of males to other males near their territory borders differ from displays directed towards the mate, but the displays have elements in common. Males call across the territory border, give “coo” calls and bow the head in display of the crest and the brightly colored bare skin behind the eye. When one male intrudes, the resident male approaches with head lowered, the bare skin fully exposed (particularly the orange skin, Folse 1974), the tail held upright and wagging from side to side. The intruder exposes his head colors, and the birds move from side to side until the intruder backs down.The resident approaches in short bursts, running and dodging, popping the wings.The resident female may join in defense as she follows the male to the border and makes “clack” sounds until the intruder is gone. In Common Cuckoos, courtship begins when the female gives a bubbling call and a male approaches, or the female approaches a calling male. The male then bobs the head or bows the body, the wings are open and drooped, and the tail is raised and fanned; in the fan posture the male may rotate the body slowly and sway from side to side.The female looks on; only rarely does the male offer her a caterpillar. In mating, the male silently follows the female and may give a piece of vegetation to her, the female opens her wings slightly, moves her tail to one side, and calls just before mating. The male mounts, then droops his wings slightly and moves his tail slowly from side to side. Copulation occurs after the female has laid an egg, and this fertilizes the next egg, which is laid 1-1/2 days later (Wiley 1981, Cramp 1985). The female areas of laying activity do not match the male territories. A male may have more than one female in his area, and a female’s area may overlap that of more than one male (Dröscher 1990, Nakamura and Miyazawa 1997). In some areas where cuckoo density is high, females are territorial and exclude other females from their laying area, even when the other females’ eggs mimic the eggs of another host species. The male sometimes aids his mate, the female in his territory, in expelling an intruding female; at other times he mates with the intruder (Riddiford 1986). Cuckoos of several species spread and raise their tails in sexual display. Secretive as cuckoos are,

glimpses of this behavior have been seen in Old World parasitic cuckoos, in New World Coccyzus cuckoos, in American Striped Cuckoo, in the anis and in Guira Cuckoo. The bold patterns of white spots on the tail are shown in breeding displays in many other Old World malkohas and New World coccyzines. Courtship feeding has been observed in broodparasitic cuckoos (Figure 8.1) including Clamator ( Jacobin Cuckoo, Levaillant’s Cuckoo, Great Spotted Cuckoo), several glossy cuckoos Chrysococcyx (Blackeared Cuckoo, Horsfield’s Bronze-cuckoo, Little Bronze-cuckoo, Shining Bronze-cuckoo, Diederik Cuckoo, Klaas’s Cuckoo, African Emerald Cuckoo), and the fruit-eating koels (Dwarf Koel, Common Koel) and Channel-billed Cuckoo where the male presents the female with fruit, and in Thickbilled Cuckoo. In these birds, courtship feeding is common. Courtship feeding is often seen in several Cacomantis cuckoos (Pallid Cuckoo, Plaintive Cuckoo, Brush Cuckoo, Fan-tailed Cuckoo), and it is seen occasionally in Cuculus cuckoos (Black Cuckoo, Red-chested Cuckoo and Common Cuckoo).

Figure 8.1. Courtship feeding by Jacobin Cuckoos Clamator jacobinus, Lochinvar National Park, Zambia.

116 The Cuckoos In courtship feeding the male captures and holds a caterpillar, then calls and attracts a female to which he gives the prey. In the brood-parasitic Diederik Cuckoo, the male has a perch on his territory where he calls with a caterpillar in his bill, and here he attracts a series of females.When he has a caterpillar he gives a call, and the female gives another call when she attracts the male, the female “caterpillar call” of “deah-deah-deah . . .” which signals courtship-feeding. The male spreads his wings and tail in display as he presents the caterpillar to her; the female repeats her own call at this time, then he mounts her as she takes the food. Females that take the caterpillar are often in the act of ovulation of the yolk into the oviduct, before the germ cell and yolk are surrounded by albumen and egg membranes (Payne 1973a), and this is the time when copulation may effect fertilization. In Diederik Cuckoos in the Eastern Cape Province of South Africa I observed courtship repeatedly, many times in an hour, day after day, as a male attracted his breeding females to his call-site. Female cuckoos may gain a significant amount of their food to form their eggs during the period of courtship feeding. Courtship feeding occurs in nesting cuckoos as well as in the brood-parasites. Courtship feeding has been seen in African yellowbills, but has not been reported in the Asian malkohas. In the New World the behavior has been seen in Dwarf Cuckoo, Yellow-billed Cuckoo, Black-billed Cuckoo, Mangrove Cuckoo, Dark-billed Cuckoo, Jamaican Lizard-cuckoo and Squirrel Cuckoo. In other cuckoos, courtship feeding has been described in couas (Red-fronted Coua, Blue Coua), in coucals (Violaceous Coucal, Pheasant Coucal, Greater Coucal, Madagascar Coucal, African Black Coucal, Lesser Coucal, Green-billed Coucal (although courtship feeding has not been seen in copulation, the male feeds the female on the nest) and White-browed Coucal), in cooperatively-breeding crotophagines (Smooth-billed Ani, Groove-billed Ani), and in the New World ground-cuckoos (Greater Roadrunner). These birds are among the more readily observed cuckoos, though courtship feeding may be wides-

pread among other species of nesting cuckoos as well.

Songs and calls Cuckoo species have unique songs and distinctive calls. Some cuckoos have melodious whistles that recall the musical song of songbirds and humans. In other cuckoos the calls are not whistled but are used in much the same social context and can be referred to as “songs” that are loud, distinctive, and used in the same behavior contexts as the songs of songbirds, in announcing their territories and attracting a mate. Cuckoo songs range from clear whistles to unmusical noises and mechanicalsounding vocalizations. The glossy cuckoos Chrysococcyx and the large Cuculus brood-parasitic cuckoos of the Old World have clear whistles that are musical to listening humans, Great Spotted Cuckoos give harsh screams; malkohas give chatters and rattles, the coucals give dove-like coos, the roadrunners give coos and rattles, the anis give hisses and glissandos, and the New World groundcuckoos make guttural sounds. Cuckoos are distinctive in their songs, especially in the Old World where several species occur together. Many cuckoos call in a position like calling doves, bringing the head forward and down until it nearly touches the breast (roadrunners, coucals, Old World groundcuckoos). Cuckoos sing to establish and maintain their territories and attract their mates.The brood-parasitic cuckoos sing repeatedly all day and even at night during the breeding season, and they sing in territorial behavior and courtship. In the nesting cuckoos, a Squirrel Cuckoo announces his territory by calling as many as 96 notes a minute for several hours through the day (Sick 1993). Females also call, as New World Yellow-billed Cuckoos do with a short version of the male call, and cuckoos often call back and forth with the inmates while on the nest, the females lacking the terminal notes. Jacobin Cuckoo females give a long call “kleeuw, kleeuw, kleeuw” like the male, but only the male follows the series with a faster series of short rising notes “kwik-kwik-kwik”, a difference in behavior that

Breeding biology and life histories 117 was useful to know when females were collected during an ovary study of the otherwise similar males and females (Payne 1973a). Females in Old World brood-parasitic Cuculus cuckoos give a bubbling call unlike the calls of the males, and females do not have loud persistent songs. Some cuckoos such as the Old World ground-cuckoos and New World ground-cuckoos such as the roadrunners signal with bill-pops, clicks and chatters, particularly in aggressive contexts. Group defense of territory against other groups in group-living anis and Guira Cuckoos involves harsh vocal calls. Squirrel Cuckoos have been said to imitate and mimic the sounds of other kinds of birds (Sick 1993). However, vocal mimicry has not been recorded in the field, and no cuckoos are known to learn their songs and calls. Vocal repertoires of more than a single song are best known in the Common Roadrunner. These birds have several distinct calls, and they are used in different situations (Whitson 1971, 1975). (1) The most commonly heard call is a “coo.” The call is a low-pitched hoot like that of a dove or an owl, a downward slurred note, given in a series of about five notes, “coo-coo-coo-coo-cooo,” by a male as he perches on a post or other elevated site. As the male calls, he lowers his head with the crest erect and the bare skin behind the eye exposed, and moves the bill away from the body with each note. (2) A short “coo” is given by the female in a series of two notes, and is not as loud, audible to about 30 m; this is also given by the male. (3) A single “coo” is a soft note heard only up to 2 m away and is given by both sexes, with a flick-bow display in courtship. (4) A “bark” is a rapid series of short notes that sound like coyote yelps and this loud call carries for up to 300 m. It is given by the female, often with the crest raised and the bare skin exposed. (5) “Growl-coo” is 3–4 notes of low frequency, given by both sexes. As the bird calls its throat bulges, it fluffs its feathers, raises its crest and usually conceals bare skin. (6) A “whirr” is a soft, low pitched sound where a “whirr” regularly comes between a series of “put-put” calls heard up to 2 m away.The male gives the call during a tail-wag display in courtship. (7) A “Whine” is a single low-

pitched call heard up to about 5 m, and is given by both sexes. The bird shakes its head from side to side as it moves the head downward. In addition to these vocal calls, roadrunners rattle their bills together in a “clack” while they produce a vocal whine in the syrinx, and the male makes a “pop” by bringing his wings together inward towards the body. In pair formation the “coo” and “bark” are usually given when birds forage apart, and the softer “growl,”“coo” and “whine” when they forage near each other. The birds become quiet while they are rearing the nestlings and more vocal when they have fledglings. Of these calls, the “coo” series is the one like the long-distance territorial song in songbirds. The call is given loudly by the male, from a conspicuous perch, and the male may call for a couple hours beginning at sunrise; the male often remains on the same perch, but he sometimes calls from more than one perch site, and males in neighboring territories alternate calls during the early morning. In the breeding season, many cuckoos call both in the day and at night, both the nesting species such as the Black-billed Cuckoo in North America and several brood-parasitic cuckoos in the Old World. In tropical regions when other birds are quiet, the cuckoos often call persistently in the hot middle of the day, like doves. They also call when rain approaches or has just passed, and for this behavior the cuckoos are known as “rain birds.” “Brain-fever” cuckoos are named for their persistent calls, as in Pallid Cuckoo in Australia, Black Cuckoo in Africa, and Plaintive Cuckoo, Large Hawk-cuckoo and Common Hawk-cuckoo in Asia.The name “Brain-fever” was used for cuckoos in India and Burma as they called all night; in Uganda an administrator was petitioned to send his collectors to shoot these birds as they disturbed the residents’ sleep ( Jackson 1938). Some cuckoos including Asian hawk-cuckoos (Large Hawkcuckoo, Common Hawk-cuckoo) even scream “Brain-fever!” Night sounds are also given by New World nesting cuckoos such as the Black- and Yellow-billed Cuckoos. In Finland an Oriental Cuckoo called day and night. When its song was recorded and played in the field at midnight, the

118 The Cuckoos cuckoo rapidly flew to the playback and was captured in a mist net in conditions so dark that its plumage color could not be seen (Lindholm 2001). These cuckoo songs appear to be as active in advertising a territory at night as in the daytime. The origins of many scientific names of birds have been traced by Jobling (1991). Several cuckoo genera were named for their calls, Cuculus itself for the onomatopoeic “cuckoo” call of Common Cuckoo C. canorus, and descriptions and interpretations of the calls (Clamator, the shouting cuckoos; Cacomantis (Gr. kakos, evil, ill-boding; mantis, prophet) refers to the “rain bird” role in folklore, as these birds are said to predict ill fortune and bad weather; its synonym “Penthoceryx” (Gr. penthos, misfortune, grief; kerux, a herald) indicates the same; “Caliechthrus” (Gr. kaleo, to call; ekhthros, odious) the loud monotonous calls, which according to tradition foretell bad weather and ill fortune; Coccyzus derives from Gr. kokkuzo, to cry “Cuckoo”; Hyetornis (Gr. huetos, rain; ornis, bird) refers to the calls of the local rain-bird; Coua is a Malagasy name for these birds that repeat the calls transliterated as “coua”; and Guira derives from güirá , a Paraguayan Indian name for a bird named after its calls.

Mating systems Most nesting cuckoos are socially monogamous: a male and a female live together during the breeding season. Both sexes rear the young.The cooperatively breeding anis appear to maintain distinct social pairs of a male and a female. In a few cuckoos there have been molecular genetic studies to test whether the members of a social pair are monogamous or mate outside the pair bond as well. Mating systems of the brood-parasitic cuckoos are not well known: male and female show no strong pair bond. A male Diederik Cuckoo regularly calls in a bush and attracts a series of females; he feeds each with caterpillars and courts them one by one over a day or several days (Payne 1973a), so there appears to be no special pair bond. In one case a male Diederik Cuckoo had two females, one that spent her time in a colony of Fire-crowned Bishops Euplectes hordeaceus, the other in a colony of Black-headed Weaver Ploceus melanocephalus

(Verheyen 1953). Males occasionally feed a fledged young cuckoo and this behavior may be misdirected courtship feeding. Adult brood-parasitic cuckoos do not feed the young while the nestling cuckoos are still in the hosts’ nest. Exceptions to the lack of a pair bond in brood-parasitic cuckoos are seen in cases when male and female act together, as when the male draws the host off the nest while the female slips onto the nest to lay her egg, in crested Jacobin Cuckoo, Common Koel and Channelbilled Cuckoo. In Common Cuckoos in Britain and Japan, the mating system was exposed by comparing nuclear microsatellite genetic profiles of adults and nestlings. In both areas, cuckoo mothers were identified by comparing their profiles with those of cuckoo chicks in the nests of the host species. Cuckoo mothers were also identified by comparing mitochondrial genes of adult females and chicks, as these genes are transmitted and inherited in a matrilineal pattern. The young of each female had the same father, and the results show that the female was monogamous. One male was polygamous and was father of the offspring of two females. The results indicate that each female prefers to lay in the nests of the host species that fostered her, but she does not preferentially mate with a male that was reared by the same foster species ( Jones et al. 1997, Marchetti et al. 1998, Gibbs et al. 2000). Cuckoos do not learn or mimic the songs of their foster species, and they do not have a vocal signal that guides a female to mate with a male that was reared by her own foster species. In Great Spotted Cuckoos a female sometimes lays more than one egg in a crow’s nest, and more than one female cuckoo sometimes lays in a nest. Microsatellite markers reveal some cases of monogamy and other cases of a cuckoo mating with more than one partner. No mitochondrial markers were included in the study, and it is unknown whether these polygamous cases were a female mating with more than one male, a male with more than one female, or both (Martínez et al. 1998a,b). In contrast to the brood-parasitic cuckoos, males and females in the brood-parasitic finches Vidua have a common genetic interest in finding a mate that was reared by the same foster species. Each

Breeding biology and life histories 119 species of indigobird Vidua mimics the mouth pattern of its host species, and the nestling indigobird has a mouth color and pattern like that of the host nestlings with which it is normally reared in the nest (Payne 1973, Payne and Payne 1994). Both male and female form a behaviorally imprinted association with their foster species. Males mimic the songs of their foster species, which they learn from their foster parents. Females are attracted to the songs of male indigobirds that mimic their own foster species. Moreover, the females reared by an alternate foster species parasitize their own foster species and not the nests of their normal host species where the nestlings would have the mouth pattern of the females’ own offspring, even when nests of the normal host species are available. The inclination of both male and female indigobirds to a common foster species indicates a common genetic interest in their offspring, as both sexes contribute nuclear genes that determine mouth mimicry in their young, and both benefit when their mate has the same developmental experience and genetic traits for nestling mimicry and the eggs are laid in the nest of the same species that had reared the adults (Payne et al. 1998, 2000, Sorenson et al. 2003a). Most breeding cuckoos are socially monogamous.The partners of a pair are often together, and both male and female incubate the clutch and rear the brood. Molecular genetic studies have been carried out with cooperatively breeding New World anis and with Guira Cuckoos. Early results in Guira Cuckoos suggest that nestmates in a communal nest are more likely to be half-sibs rather than full sibs, indicating that a parent had mated with more than one partner. The genetic profiles of nestmates in one nest, revealed that a female mated with more than one male and a male mated with more than one female (Quinn et al. 1994). Fieldwork continues in social behavior and molecular genetics of families in these cuckoos. In the South American Dwarf Cuckoo Coccycua pumila and Dark-billed Cuckoo Coccyzus melacoryphus occasionally two females lay in the same nest. The birds may be polygynous with two females mated to the same male, or they may be facultative brood parasites (Ralph 1975, Sick 1993). In North

America, Black-billed and Yellow-billed Cuckoos occasionally lay in the nests of other species such as songbirds and doves, and more often they lay in each other’s nests (Darwin 1875, Bendire 1895, Nickell 1954a, Wiens 1965, Nolan and Thompson 1975, Peck and James 1983, Fleischer et al. 1985). In these cases it is unlikely that interspecific promiscuous sex was involved. Molecular genetic studies and behavioral observations of marked cuckoos are needed to test whether the cuckoo eggs and nestlings in these nests are related at the half-sib level, and whether one male is the mate of both females. Perhaps a female lays in the nest of another pair when her own nest has been disturbed. In the nests where these cuckoo eggs appear, the nesting species sometimes rears the young cuckoo (Darwin 1875, Nickell 1954b); more often the outcome is unknown. Most coucals that have been watched in the field are monogamous, living in pairs (Baker 1934, Frauca 1967, Dhindsa and Toor 1981, Taplin and Beurteaux 1992, Natarajan 1997, Hustler 1998). Males often carry out the larger share of parental care, with both sexes incubating and caring for the young. Behaviors that suggest a strong social pair bond in coucals includes courtship feeding by the male and song duets by the pair. These behaviors are more common in monogamous birds with a strong pair bond than in birds with other mating systems (Lack 1968, Payne 1971, Kunkel 1974, Farabaugh 1982). On the basis of the sex roles in parental care, some coucals are suspected to be polyandrous. In polyandrous birds a female mates with more than one male and in these birds often the male takes a more active and conspicuous role than the female in incubation and care of the young. Polyandry has been seen once in the African Black Coucal, when a female had three males and mated with each male in turn, while each male held a separate territory within her large domain.The males built the nests, the female laid the eggs, then each male incubated the clutch in a nest and brooded and fed the young. The female took no active role in parental care. Only two females were observed, and only one had more than one mate (Vernon 1971). Reversed sexual size dimorphism with females larger than males coincides with clades or lineages

120 The Cuckoos of coucals, rather than with widespread reversal of sex roles or polyandry as suggested by Ligon (1993, 1999) and Andersson (1995). As in other birds, the female coucals have only one ovary and oviduct. Male coucals often have only one large testis (Rand 1933, 1936, Chapin 1939, Centropus milo in UWBM) and when two testes are present they vary in whether the left or right testis is larger even within a species (Mayr 1937, Mayr and Rand 1937, Rand 1942a), while the testes are extremely asymmetric in some other cuckoos such as the cooperatively breeding anis (McNeil 1968). Explanations of the origin and adaptation to polyandry in coucals in terms of their asymmetric testes (Ligon 1997, 1999) are not well supported, insofar as field studies show that coucals are generally not polyandrous.The idea that nocturnal incubation by the male in cuckoos led to increased male parental care and to relief of the female from nesting duties in cuckoos, suggested by Vehrencamp (1982a, 1985) and Ligon (1993, 1999), could be explored by determining the incubation roles of males in other kinds of polyandrous birds. Does staggered asynchronous hatching make it likely that a single parent can care for a brood? Again, there is no evidence that asynchronous hatching which is common in other cuckoos is associated with polyandrous mating. African Black Coucals are seasonally migratory in parts of their range, and Andersson (1995) suggested that migratory behavior may have led to polyandry, yet other

migratory cuckoos are not polyandrous.Additional fieldwork is called for on breeding behavior of these cuckoos.

Nests and nestbuilding Nest structure with most cuckoos is simple. The cuckoos build an open platform nest, using sticks as the main nest material, and they line the nest with finer sticks (Figure 8.2).These nests are often flimsy and the cup that holds the eggs is shallow. Coucals build a covered nest, sometimes called a globular or domed or thatched nest. These nests are usually built of grass or other soft vegetation, or sticks thatched together with the stems of grass in which the nest is concealed. The sides are built upward and form a roof. The nest has an entrance on one side, and in some the entrance has a covered tunnel that leads between the nest chamber and the surrounding vegetation. A few cuckoo species have a variable nest structure, sometimes open and sometimes covered. Pheasant Coucal Centropus phasianinus, Greater Coucal C. sinensis and Madagascar Coucal C. toulou sometimes build an open nest rather than a covered nest, and only an open nest has been described for Bay Coucal C. celebensis in a group which otherwise have covered nests. Also, White-browed Coucal C. superciliosus builds a bower-like nest when it is built in dense overhanging vegetation. Senegal Coucal C. senegalensis usually build a covered nest.

Figure 8.2. Black-billed Cuckoo and nest, Ann Arbor, Michigan: a, adult on nest; b, nest with 3 eggs.

Breeding biology and life histories 121 However, one pair laid in an open nest that they built in a dense spiny citrus tree where the canopy overlaid the unroofed nest when the eggs were laid—we saw this at Bakary Sambou, near Brikama, The Gambia, in early October 1999. In some cuckoos that usually build an open nest, the yellowbills Ceuthmochares and Blue Coua Coua caerulea sometimes build a covered nest.Timing may determine when the nest is completed and the eggs are laid, with some building continuing after the first eggs are laid, as in the open nest of the coucals. Or the nest structure may vary with the nesting substrate, as in the coucals. Some cuckoos continue to bring nesting material to the nest, especially green leaves, after the nest has been built, during incubation and after the young have hatched.This behavior has been seen in coucals, anis, Guira Cuckoo, roadrunners and the New World ground-cuckoos. It is unknown whether the leaves have an antibiotic effect that keeps bacteria from invading the eggs or young, or a strong odor that covers the odor of the brood and avoids detection by a predator.

Nesting pairs Most cuckoo species that build their own nests and rear their young appear to be monogamous, the birds appearing in pairs and two birds caring for the eggs and young in the nest.The nesting cuckoo that has been observed in most detail, Greater Roadrunner, shows behavior that appears to be typical of a monogamous bird. The male brings food—mice, small birds, snakes and lizards—to his mate before mating and during copulation. The pair copulates at the place where they will build the nest.They continue to copulate after their clutch is complete, and copulation appears to have a social function in bonding the pair as well as a fertilizing function when the female is ovulating. The nest is usually built in isolated thickets of small trees such as mesquite and bushes, in an area near open ground or short grass where the birds display and forage. The nest site is 1–3 m above the ground, occasionally higher than this, in the crotch of a tree or resting along a horizontal branch, near the center of the bush or tree. Nests are built mainly by the

female. The male brings building material in the form of thorny twigs and branches to the nest site, and gives them to the female when she gives a “whine” call. She incorporates them into the nest, laying the larger sticks around the edge and smaller sticks near the center of the nest. The pair may begin several nests and desert a site after building for a few minutes (Whitson 1975). In the site selected for the active nest, she builds the platform of sticks, then adds a lining of finer nest material such as leaves, grass, mesquite pods, snakeskin, and dry flakes of cattle and horse manure as well as feathers.The nest is a flattened cup, about 30 cm in outer diameter, 5–10 cm in inside depth and 15–20 cm in outside depth (Sutton 1940, Hughes 1996b). Some pairs continue to build during incubation and the nest may grow in height as the nestlings hatch and grow and require a larger and more protected living space (Calder 1976). Roadrunners lay a clutch of smooth white eggs, usually 3–6 (Meinzer 1993). Occasional sets of 9–13 eggs may be mixed clutches in which more than one female laid in the same nest (Sutton 1940), although the number of eggs a single female lays can vary. The clutch is often larger after good summer rains when food is abundant.The eggs are usually laid on alternate days but also at irregular intervals (1 to 4 days or as long as 9 days), and the birds begin to incubate before the complete clutch has been laid.The eggs are covered continuously, as the incubating parent remains on the nest until the other parent arrives to relieve it. Incubation that effectively transfers heat begins only after the parent has covered the eggs, and as some eggs are laid after incubation has begun, the eggs hatch asynchronously on a staggered schedule over several days. Roadrunner eggs hatch in about 18 days.The young in a brood are of different sizes with age differences as great as 7 days between the largest and smallest siblings.When the young bird hatches, the parent carries away the eggshell, then she breaks and eats it (Calder 1967a, Muller 1971, Smith 1981). The last-laid eggs are often abandoned after the first-hatched young have left the nest (Folse 1974). The last young may remain in the nest up to a week after the other nestlings have fledged. Success of the brood varies with feeding conditions,

122 The Cuckoos where parents are able to rear all their young when food is abundant. The young remain in their nest while both parents bring food. Both parents brood their young, nearly continuously for the first four days, and at cool times later.The parents bring intact food items such as lizards and snakes and the only treatment they give is to beat the vertebrate prey into a flexible form, and for grasshoppers they remove the hind legs (Meinzer 1993). The adult eats the nestlings’ fecal sac.Adults eat their own young nestlings when the young are unresponsive or weak, or feed these young to the larger and stronger nestlings when food is short. Nestlings may evict their younger siblings from the nest (Hughes 1996b), a behavior like that in the brood-parasitic cuckoos. Nestling roadrunners give a loud vocal “churr” and they rattle the bill when disturbed and excrete a blackish foul-smelling liquid reminiscent of disturbed snakes. When the young have grown, the parents give soft “coo” calls to the young to leave the nest (Calder 1967).The young fledge in 17–19 days, or as late as 25 days, or earlier if they are disturbed in the nest. Roadrunner young at fledging are only 50% of average adult weight, and young have been observed foraging with their parents when still only 70% of adult size. The adults lead their young far from the nest a few days after they fledge. Fledged young follow their parents to feeding areas, quivering their wings as they beg for food (Meinzer 1993) until they can feed themselves, with independence gradually developing about 30–40 days after they leave the nest (Whitson 1975). Roadrunner nests are spotted by watching and following an adult as it perches or runs over the ground carrying a snake, lizard or other large item of prey in its bill (Figure 8.3).The bird makes a beeline for the nest shrub, then hops up to the nest. Laura and I watched a roadrunner in southern Arizona. It crossed the road with a lizard in its bill, ran to the base of a mesquite, then ran up the sloping trunk and branches directly to a nest about 5 m above the ground, and fed the nearly-grown young in the nest. The female roadrunner sometimes lays a second clutch while her fledglings are being fed by the

Figure 8.3. Greater Roadrunner Geococcyx californianus with food for its young, Joshua Tree National Monument, California.

male, and this biparental care allows the pair to hatch and raise two broods. In Texas one pair had three nests in succession, the first in April, the second in June and the third in August, and all three broods fledged.Another pair began a second nest in July while the pair were still feeding the fledglings from the earlier nest. In some renestings the same nest is used, in others a new nest is built (Meinzer 1993). Parental behavior has been studied using radio-telemetry. The male incubates at night, and the pair takes turns during the day with the female taking two long spells on the nest. Males that incubate at night maintain a body temperature of about 2°C lower than in the daytime. Females and

Breeding biology and life histories 123 non-incubating males drop their body temperature 5°–8°C at night and save about 36% of their energy requirements during this period. Breeding males are more robust with fat deposits and heavier body mass than nonbreeding males, and breeding female roadrunners recover their prebreeding body mass while their males work the night shift (Vehrencamp 1982a). Roadrunners may be typical of nesting cuckoos in their nesting, although it is hard to know and most cuckoos have not been observed in detail. In the features that may be widespread among nesting cuckoos, both the male and the female incubate, feed the young in the nest, and tend the fledglings. In Black-billed Cuckoos a pair alternate their parental care during incubation with each member on the nest for about two hours. In most coucals both parents care for the young, and in some the males may take a greater part than females. Also, courtship feeding is widespread in nesting cuckoos; Black-billed Cuckoos and coucals pass food from one mate to the other, a behavior that provides the female with extra food to develop her eggs. In nesting cuckoos, courtship feeding may give the female a cue to assess the male as a provider of parental care before she decides to mate with him. Couas are often seen in pairs. The birds build a shallow nest of sticks, and both parents rear the young. Courtship feeding has been seen in two coua species, Coua reynaudii and C. caerulea (Goodman et al. 1997). The breeding biology of Australian Pheasant Coucals and other coucals differs from the biology of many altricial tropical land birds.The coucals lay large clutches, have a short nesting cycle, the nestlings grow rapidly, and the juveniles mature early. Some other tropical cuckoos also have these traits. Many other tropical bird families have small clutches, the nestlings grow slowly and the young mature slowly as juveniles. Coucal nesting success is high (Taplin and Beurteaux 1992). The adults add green leaves to the nest before laying and continue to add leaves through incubation. Nestling coucals have curious hair-like natal down formed by a long keratinized sheath of the growing contour feather. Nestlings also have prominent marks and papillae on the palate and sometimes these have contrasting

colors.The young have a prominent egg-tooth as in other cuckoos, and they emit a foul-smelling liquid excreta from the cloaca when nestlings are disturbed in the nest (Frauca 1967).

Eggs Cuckoo eggs are remarkable in the brood-parasites, as they often match the color and pattern of the eggs of their hosts (Davies 2000). In contrast, most nesting cuckoos have plain white eggs.The general features of egg size, shape, and thickness are not unusual in cuckoos as a group, but the eggs of brood-parasitic cuckoos are special in these features, which appear to be adaptations to their breeding behavior. Some features of egg size and shape in cuckoos, however, appear to be more directly accounted for by the evolutionary relationships of these birds than by the special demands of their breeding behavior.

Egg size and shape The eggs of nesting cuckoos are similar in size to that of other altricial birds of their body size, but the eggs of brood-parasitic cuckoos tend to be small (Darwin 1875, Payne 1974;Table 4.3, Figure 8.4). The small eggs of brood-parasitic cuckoos are similar in size to the host eggs. Small eggs may allow a female parasitic cuckoo to lay more eggs in a season, for reasons of economy, and the eggs may go undetected in the nest if they resemble the host eggs in size as well as in color and pattern.Also, small eggs may allow the female to find and lay her eggs in a larger number of nests insofar as small host species are more abundant than larger hosts the size of the cuckoo (Payne 1974). However, at least in Britain, small hosts are not parasitized proportionately more often than larger hosts (Soler et al. 1999). Common Cuckoo egg size does not vary with host egg size (Latter 1902, 1905). The exceptions were the large cuckoo eggs in nests of the large Corn Bunting Emberiza calandra, but this host accounted for only 27 of 11,870 cuckoo eggs, or 0.002 percent of the sample (Moksnes and Røskaft 1995).The trend for small eggs is unique to cuckoos among all broodparasitic birds. The other brood-parasites (the

124 The Cuckoos

Figure 8.4. Egg size and female body size in cuckoos. Egg mass (g) is estimated from the equation, g = 1.08 ⫻ 0.512 ⫻ 1 ⫻ d2, where 1.08 is egg density (g/mm3), 0.512 is a geometric factor, l is egg length (mm) and d is egg width (mm) (Payne 1977c).

honeyguides, the cowbirds and the African viduid finches) do not have small eggs for their body size. The cuckoos are several times larger in body size than most of their host species, whereas the other brood-parasitic birds are not as disproportionately large (Payne 1977b,c, 1989). In nesting cuckoos, the eggs in coucals are smaller than in couas, malkohas and large New World roadrunners and ground-cuckoos.The communal-nesting anis and Guira Cuckoos have large eggs for their body size. In addition, the eggs of these crotophagine cuckoos are remarkably variable in size within a local population, perhaps owing to differences in physical condition of laying females within a social group, perhaps with nestlings hatched from large eggs having a size advantage within a crowded communal brood. Nevertheless, in Guira Cuckoos the egg characteristics of size and markings vary as much within a female as between

females and there is no suggestion that females can recognize each other’s eggs (Cariello et al., 2002, 2004). Cuckoos in comparison with other birds have rounded eggs, although less round than in parrots, owls, trogons and coraciiforms (Rahn and Paganelli (1988). Among the cuckoo species, large eggs tend to be rounder than small eggs (Figure 8.5). The trend is the reverse of the trend in other orders of birds in the same size range. Brood-parasitic cuckoos lay elongated eggs, and nesting cuckoos lay round eggs. And the largest brood-parasite eggs are more elongate than the eggs of nesting cuckoos of the same egg size, suggesting that parasitic cuckoos tend to have more elongate eggs than nesting cuckoos, independently of their size. Eggs of the brood-parasitic crested cuckoos Clamator are more rounded than those of other brood-parasitic cuckoos of the same size. Brooker

Breeding biology and life histories 125

Figure 8.5. Egg shape in cuckoos. The shape (elongation) is the ratio of length to width; large eggs tend to be less elongate (more round), and the eggs of brood parasites tend to be more elongate (less round) than the eggs of nesting cuckoos of the same body size.

126 The Cuckoos and Brooker (1991) suggested that the round shape is an adaptation to avoid damage when more than one cuckoo egg is laid in a host nest. Great Spotted Cuckoo Clamator glandarius lays while perched on the edge of a nest, and the host eggs in nests that have been parasitized are sometimes cracked. In experiments, when eggs are dropped from this height, on impact they crack the host eggs in the nest (Soler et al. 1996). A few other brood-parasitic cuckoos also lay more than one egg in a nest, particularly Common Koel Eudynamys scolopacea and Channel-billed Cuckoo Scythrops novaehollandiae, but these cuckoos do not lay rounded eggs. In addition, the egg shape in Clamator is about the same as in the malkohas, birds of the same body size that do not lay communally in a single nest. Finally, in the crotophagine anis, cooperative breeding birds in which several females lay competitively within a common nest, the eggs are not round, except in the largest species, Greater Ani, in which body size is large and egg shape is in line with the trend for large birds to lay rounded eggs. Because the prediction of rounded eggs with multiple nest parasitism does not account for egg shape in these cuckoos, it is more likely that eggs of brood-parasitic Clamator are round due to their ancestral relationship to the nesting malkohas.

Eggshell thickness, strength and ultrastructure Cuckoos as a group do not have unusually dense eggshells. In a comparison of egg structure in all avian orders, the only birds that were found to have unusually thick eggshells and eggs with high shell mass for their body size were the galliforms (Rahn and Paganelli 1989). Eggs of parasitic cuckoos may have a hard and thick shell that resists cracking when the female lays her egg from a perch above the nest.The host often attempts to remove the cuckoo egg from its nest, and the eggs of parasitic cuckoos might be structurally formed to avoid being rejected. Some hosts, mainly the larger species, remove the eggs by grasping around the eggs with their bills, while the smaller hosts with smaller bills can remove the eggs by puncturing the shells, grasping the edge and lift-

ing the eggs from the nest. A structural adaptation by which the eggs of brood-parasitic cuckoos might avoid rejection is a thick or dense eggshell that resists puncture by the hosts’ bill. Another is a round shape, which would be more difficult for the host to grasp and to puncture. In an experimental study of eggs of brood-parasitic cowbirds, which have a thicker eggshell and rounder shape than icterid species that rear their own young, the cowbird eggs that were thicker-shelled and more rounded had a greater resistance to being punctured than eggs that were thin and more elongate. In the cowbird eggs, eggshell thickness and egg shape contributed equally to puncture resistance and this structure appears adapted to their brood parasitism (Picman 1989). To test whether brood-parasitic cuckoos have eggshells that are thick and dense compared with those of nesting cuckoos, the data on eggshell thickness and mass in Schönwetter (1964) were compared with egg size in Table 4.3. Eggshell thickness and weight were closely associated; eggshell weight is used in the comparison of egg size and eggshell mass (Figure 8.6). The eggshell is not thicker or denser in brood-parasitic cuckoos than in the nesting cuckoos, when egg size or body size is taken into account.Within the region of similar egg size and body size, the parasitic cuckoos do not have a more robust eggshell. Picman and Pribil (1997) found that Cuculus have denser eggshells than Clamator, and this is not the result predicted if eggshell density were related to multiple nest parasitism, nor is the difference apparent in the data when egg size and body size are taken into account. The eggshell of Cuculus eggs is thicker than that of their nesting hosts, and the hatching cuckoos began to peck earlier in relation to hatching and made more pecks at the shell from inside the egg than did young Great Reed Warblers Acrocephalus arundinaceus (Honza et al. 2001). It remains to be tested whether the eggs of the brood-parasitic cuckoos and nesting cuckoos hatch in the same manner. The eggshell structure in brood-parasitic cuckoos differs from that in nesting cuckoos (Mikhailov 1997). Cuckoo eggshells are like those of coraciiform and piciform birds, and differ from

Breeding biology and life histories 127

Figure 8.6. Egg size and eggshell weight in nesting cuckoos and brood-parasitic cuckoos.

other orders of birds in having two outer layers in distinct and sharply separated zones (SqZ ⫹ EZ). An external zone is present in all cuckoo eggs and is absent in passerine eggs. In nesting anis, Guira Cuckoo and roadrunners there is a thick carbonate microglobular outer layer (15–40 µm) that appears chalky on the surface (Schmidt 1964, Board and Perrot 1979). The egg is covered with vaterite, a form of calcium carbonate that forms small globes. This material is resistant to crushing (Tyler 1969) and it may be an adaptation to the competition over eggs that occurs in the communal nest of anis where one female drops her eggs onto the eggs of another female. In eggs of Centropus and Coccyzus, the cuticle is thin (3–10 µm) with microglobular and granular inclusions, and the egg looks dull and non-crystalline, again with a bloom. Eggshells of some brood-parasitic cuckoos have similar globeshaped particles on the surface (Becking 1975a), and this layer may be an adaptation for parasitism in

birds that drop their eggs into a nest. The eggs of most brood-parasitic cuckoos are glossy and lack the thick layer of vaterite that characterizes ani eggs. The eggshells of Old World parasitic cuckoos have a thick outer layer that differs from eggshells of nesting cuckoos and are like those of some other birds such as the Passeriformes (Mikhailov 1997). Although these features of the eggshell of parasitic cuckoos may contribute to resistance to damage by the hosts, no functional design is obvious in the microstructure of the eggshell.

Incubation behavior, incubation period and temperature tolerance In most cuckoos both sexes incubate the eggs, transferring heat from the body of the adult to the eggs in the clutch. Heat is transferred through a brood patch, a bare area on the ventral side of the adult.The brood patch becomes vascular and thick

128 The Cuckoos in breeding Yellow-billed Cuckoos and presumably in other nesting cuckoos. Both sexes incubate and both are said to have a brood patch (Pyle 1997). The brood patch has not been examined in nesting cuckoos particularly the male, and some observations may have mistaken the ventral apterium which all cuckoos have for the specialized shortterm brood patch which occurs only when birds are nesting. There is no brood patch in broodparasitic cuckoos, though some have been reported (e.g. Thompson 1966). In fact no brood-parasitic birds of any kind are known to have a brood patch. The brood-parasitic cuckoos have a shorter incubation period than their hosts (Payne 1977b, Davies 2000). The incubation period is short in many nesting cuckoos of similar body size as the brood-parasitic cuckoos as well. For this reason it is uncertain whether the incubation period has been selected to be short in the brood parasites. It is likely to be so, if only because other brood parasitic

birds such as cowbirds Icteridae and parasitic finches Viduidae also have shorter incubation periods than their host species and have shorter periods than that of related species of nesting icterids and nesting estrildid finches (Payne 1977b). The egg of a parasitic cuckoo is held for a day in the oviduct of the female before she lays.This head start in development of the embryo contributes to shortening of the incubation period of these brood parasites (retention of the egg is unknown in other broodparasitic birds, Payne 1965, 1973a, 1977c, 1989). When a female cuckoo lays an egg, the embryo is developed to a 24-hour stage where the primitive streak can readily be seen on the yolk surface within the egg (Perrins 1967, Vernon 1970a, Liversidge 1971, Payne 1973a). Davidson (1886) appears to have been the first to note that a cuckoo egg laid in a parasitized nest is more developed than the host eggs. Retaining the egg in the oviduct for a day is also useful when a female cuckoo finds that

Figure 8.7. Egg size (g) and incubation period (days) in cuckoos.

Breeding biology and life histories 129 a target nest is lost to a predator, as she then has time to find another nest. Brood-parasitic cuckoos have short incubation periods, with 12 days in the smaller species and 15 in the larger species (Figure 8.7), a general trend in birds with about the same body size (Calder 1984). Nesting cuckoos also have short incubation periods when compared with other nesting altricial birds. There is no clear difference in incubation period between brood-parasitic and nesting cuckoos, when the longer time for incubation of larger eggs is taken into account.The brood-parasitic Clamator have an incubation period about the same as the small Cuculinae (Coccyzus species). Most of the variation in incubation period in cuckoo species is due to the size of the egg and not to whether the cuckoo is a brood parasite. In addition to a short incubation period, the developing cuckoo embryo is more tolerant of cooling than the embryo of most birds. In Australia,

an egg of Horsfield’s Bronze-cuckoo was taken from a nest along with eggs of the host Malurus sp. fairywren.The cuckoo embryo was still alive in the egg 24 hours later, while the fairy-wren embryos were dead (Serventy and Whittell 1976). Tolerance of developing cuckoos to cool temperatures continues into the nestling period, as discussed below.

Development of the young Cuckoos are altricial birds. Nesting cuckoos remain in the nest while they grow.They depend on their parents or foster parents to provide food, and they grow rapidly. Nestling arboreal cuckoo young (Little Cuckoo,Yellow-billed Cuckoo, Black-billed Cuckoo, Squirrel Cuckoo, Smooth-billed Ani, Groove-billed Ani) grow rapidly and leave the nest in as little as 10 days, and can climb about when they are disturbed and fledge earlier if at risk from predation. The larger ground cuckoos including

Figure 8.8. Egg size (g) and nestling period (days) in cuckoos.

130 The Cuckoos Coral-billed Ground-cuckoo, Pheasant Coucal, White-browed Coucal and Greater Roadrunner fledge in 17–20 days, a longer time than in the smaller nesting cuckoos. The nestling period of brood-parasitic cuckoos is longer than that of nesting cuckoos (Figure 8.8). Nestlings of the small brood-parasitic cuckoos (Chrysococcyx basalis, C. lucidus, C. caprius, C. klaas, C. cupreus,Tapera naevia) take 18–20 days to fledge, Common Cuckoo take 18–20 days, Common Koel take 20–28 days, Great Spotted Cuckoo take 22–26 days, and Channel-billed Cuckoos take as long as 24 days to fledge. The nestling period of small brood-parasitic cuckoos is twice as long as the nesting cuckoos of the same body size, and they often monopolize the parental care of their foster parents for a few more weeks after they leave the nest. In the brood-parasitic Jacobin Cuckoo and Levaillant’s Cuckoo the nestling period is shorter than in Great Spotted Cuckoo or other parasitic cuckoos. The nestling period in these Clamator is

the same as that observed in the nesting cuckoos, notably the New World malkohas (breeding information is less well known for the Old World malkohas). The short nestling period in Clamator may be explained by their choice of hosts and the advantage to fledge together with the host young in the mixed-species brood. Nestling Clamator often grow up together in the nest with their foster young, rather than evicting the host eggs and young like most other brood-parasitic cuckoos do. The babbler host young of Jacobin and Levaillant’s cuckoos fledge early from the nest (Gaston 1976, 1977, Hustler 1977b), and in contrast the crow host young of Great Spotted Cuckoo remain in the nest for more than three weeks (Cramp et al. 1994). In these cuckoos it appears that the nestling period is adjusted within the brood to match the nestling period of their host young, and by fledging together in the brood the young cuckoos are likely to be fed and receive parental care after the foster parent leads its brood away from the nest.

Table 8.1 Size at fledging in cuckoos in relation to parental behavior. Species Chrysococcyx basalis Chrysococcyx caprius Chrysococcyx lucidus Clamator glandarius Clamator jacobinus Cuculus canorus Cuculus clamosus Cuculus solitarius Eudynamys scolopacea Pachycoccyx audeberti Centropus phasianinus Centropus senegalensis Centropus sinensis Coccyzus americanus Coccyzus erythropthalmus Crotophaga ani Dasylophus superciliosus Geococcyx californianus a Fledging weight/adult weight. n, nesting cuckoos, reared by their own parents. p, brood parasite.

Weight (g)a 20.9/23 (91%) 35/38 (92%) 22/25 (88%) 110/134 (82%) 56/80 (70%) 90/112 (80%) 62/90 (69%) 60/76 (79%) 125/225 (56%) 83/110 (75%) 123/302 (41%) 125/169 (74%) 160/310 (52%) 38/60 (63%) 23.8/50 (48%) 100 (35%) 60/117 (51%) 150/320 (47%)

Breeding behavior (p, n) p p p p p p p p p p n n n n n n n n

Breeding biology and life histories 131 Development of young cuckoos can also be compared in terms of their size at fledging. Growth rates in altricial birds vary with their altriciality and precociality (with higher growth rates in altricial birds which complete more of their development in the nest) and tend to be lower for birds of larger adult body size (Ricklefs 1973, Starck and Ricklefs 1998). Nestling brood-parasitic cuckoos continue to grow until they nearly reach adult body size, then they leave the nest. Nesting cuckoos such as roadrunners and anis fledge before they are fully grown. In Table 9, the fledging weights, adult weights and breeding behaviors are extracted from the cuckoo species accounts.Young of the brood-parasitic cuckoos stay in the nest until they are well grown, with most species waiting until they are at least 75% of adult body weight. The nesting cuckoos for which there is data fledge when they are only half grown. Other cuckoos have not been studied. In some species disturbance of the nest to measure the birds can cause the nestlings to fledge prematurely; as in

young Guira Cuckoos that are more than half the adult size when they normally fledge, but the young leave early when the nest is disturbed (R. Macedo, in litt.). Nesting cuckoos are unusual among altricial birds in leaving the nest long before they are able to fly. This is the pattern in most nesting cuckoos, including birds for which no weights at leaving the nest are available as in Piaya cayana and Crotophaga sulcirostris. The brood-parasitic cuckoos are unusual among the cuckoos in remaining in the nest until they are nearly grown. Their foster parents continue to care for the young parasitic cuckoos long after they leave the nest, well beyond the period of post-fledging care in the nesting cuckoos. Even in the small broodparasitic Shining Bronze-cuckoo, the foster Western Thornbills continue to feed their fledged cuckoo for as long as 42 days after fledging (Payne and Payne 1998a), well past the time when thornbill young would be independent and the parents could nest again (Nicholls et al. 2000).

9 Cooperative breeding

Cooperative breeding birds live in social groups. In the group some birds, the “helpers”, help rear the young of other individuals, the breeders. Helpers may provide relief to the breeders by taking part in the care of a brood, and the helpers may increase the number of brood young that survive. This behavior is curious. The helpers, by not breeding themselves, may not be behaving in their own best interest. This altruistic behavior would under most conditions be selected against, if only because other individuals would leave more offspring and the helpers themselves would fail to reproduce (Stacey and Koenig 1990, Ligon 1999). Nevertheless the New World anis and Guira Cuckoo are among the most social of all cooperative breeding birds. Two ideas explain how cooperative breeding is a successful behavior. First, cooperative breeding may relate to the benefits of living in a social group. Group-living birds can find more food and defend their brood against a predator (Payne et al. 1985, 1988, Rowley and Russell 1997). Group-living birds can hold an area against other social groups, as when the resident group is larger than the competing group, and the helpers not only gain resources in the present but also take over this extra land themselves in the future (Woolfenden 1975, Cockburn 1998, Kokko et al. 2000). Another way the helpers can aid their own survival and future breeding success is in gaining experience in parental care (Komdeur 1996). Finally, a “helper” may also be a breeder at the same time, as when a female lays in the nest of the older breeding female (Brown 1987, Cockburn 1998). These complex societies are potential arenas for reproductive conflict: when more than one adult female is in a

group, there is a potential opportunity for all to breed and for a socially dominant female to coerce others to care for her own offspring (Reeve et al. 1998, Johnstone et al. 1999). Second, the helpers are often closely related to other birds in their group. They may be the older offspring from an earlier brood of the parents or they may be otherwise closely related, such as brothers helping their sisters. The more closely related the helpers are to the breeders and to the young they care for, the more likely they are to share the same genes. By providing help to relatives the helpers are indirectly increasing the success of their common genes in the next generation (Hamilton 1964). This is Hamilton’s theory of inclusive fitness. In contrast to early predictions from the theory of inclusive fitness and the evolution of behavior that favors relatives, ideas that were developed in studies of social insects, recent studies of social insects in which several females breed in a nest show a low degree of genetic relatedness among the breeders (Hamilton 1972, 1996, Queller et al. 1988, Strassmann et al. 1991, 1992). Cooperative breeding in certain birds is associated with genetic relatedness and kin discrimination that is based on social experience between siblings reared in the same social group (Emlen and Wrege 1988, 1989, Stacey and Koenig 1990, Emlen 1997). Cooperative breeding also occurs in birds where helpers are not closely related to the brood or parents of the brood; and in a social group where not all birds are closely related, the helpers are as likely to care for young to which they are not closely related as to young to which they are (Payne et al. 1985, Dunn et al. 1995, Dunn and Cockburn 1996,

Cooperative breeding 133 Magrath and Whittingham 1997, Wright et al. 1999, Finn and Hughes 2001). Cooperative behavior involves three behaviors of the “helpers.” First, the birds remain in a social group rather than dispersing to their own territories. Second, they do not breed themselves. Third, they care for the dependent young of other individuals (Wiley and Rabenold 1984, Mulder and Langmore 1993, Green et al. 1995, Emlen 1995, 1997). These behaviors are not rules that all cooperatively-living birds follow, and the cuckoos are creative in the way they breed cooperatively. In the cooperatively breeding cuckoos, the anis and Guira Cuckoo, more than one female often lays in the nest (Vehrencamp 1977, 1978, Loflin 1983, Vehrencamp et al. 1986, 1988, Macedo 1992, 1994, Macedo and Bianchi 1997a,b). Field studies are now in progress to determine the social interactions and reproductive success of individual females in the cooperative-breeding cuckoos and may give answers to the question of whether group living provides opportunities for hopeful breeders, or control over the lives of the younger females by socially dominant females in the group, or both kinds of reproductive success. 1. Guira Cuckoos live in open habitats in South America.They often lay in the nests of other birds. Because their eggs appeared in the nests of other birds, not only passerines but also such unlikely birds as caracaras, there was doubt that these cuckoos built their own nests, although nesting behavior had been reported in the field (e.g. Euler 1867). When the guiras were seen to build their own nests in captivity in the Copenhagen Zoo, the doubters conceded that the birds were not obligate nest parasites (Leverkühn 1894). Guiras sometimes nest in the same bush with Smooth-billed Anis C. ani (Euler 1867). Azara (1809: 23) reported that both species often lay in the same nest: “J’ai vu plusieurs de ce nids dans lesquels étaient des oeufs des deux espèces”.Azara’s observations have often been interpreted as two species sharing parental care at a common nest (e.g. Sclater and Hudson 1889, Sick 1993) but no one else appears to have seen the cuckoos do this. Although Guira Cuckoos breed together in a communal nest, there is much conflict and com-

petition among the breeding adults. A breeding group has as many as ten females and several females lay in a nest where the common clutch has as many as 20 eggs. Adults perch on the nest, remove an egg from the nest and drop the egg on the ground, and they also fly off with the egg and bury other eggs in the nest. In Brazil, 32% of all eggs laid were lost or disappeared, and 76% of these eggs were found on the ground below the nest, tossed out by other adult guiras during laying and incubation. Other eggs are ejected shortly before hatching, and chicks are sometimes removed and killed by infanticidal adults (Macedo 1992, 1994, Quinn et al. 1994, Macedo and Bianchi 1997). In spite of the adults’ conflict over breeding and interference with each other’s nesting success, the adults also “cooperate” in incubating the clutch, feeding the brood and guarding them from predators. Usually all young in a nest hatch on the same day. At several nests all the adults attended the nest and cared for the brood, and the young were fed equitably except in the larger broods, where some young were underfed. Nests of the larger breeding groups did not fledge more young, and the number of young fledged per adult did not differ among the small, medium and large breeding groups. The probability of fledging at least one young did not vary with the number of adults. In groups with more adults, more eggs were laid but so were the number of eggs that were lost. After the brood hatches, an adult remains in attendance in the nest tree while other adults forage away from the nest. When a threat approaches the nest, the attendant gives loud staccato calls, and the other group members converge to the nest in defense (Macedo 1992, 1994). The genetic markers of young Guira Cuckoos were compared with the markers of the adults. In this study, chicks in a nest with two or more breeding pairs were often the offspring of more than one pair of adults. The chicks were usually half-siblings but not full-siblings, indicating that the adults were not sexually monogamous. Two or more females accounted for more than half the young in a communal nest. Although a single female was not responsible for most of the young in a nest, a dominant female may leave more than her share of

134 The Cuckoos young. Adults in a breeding group were sometimes related and sometimes not related to each other, and there appeared to be no strong kinship among members of a group (Quinn et al. 1994). The eggs of a female are as variable as the eggs of different individual females in their size and appearance as determined by egg proteins (Cariello et al. 2002, in press). Guira Cuckoos are difficult to study: they nest high in a tree and the birds are hard to capture and mark as individuals. Most inferences about social relationships within a group has come from molecular genetic studies rather than from observation of marked birds. 2. Groove-billed Anis breed in groups or pairs in open habitats with scattered trees. Most groups have two or three pairs. Within the social group each female has a mate with which she spends time in a social bond. Females lay eggs in a single communal nest. In the breeding season all members of a group use the territory; in the dry season the birds move into the areas of other groups, or abandon their territories and move to the edge of a forest. Anis in social groups with a larger number of adult birds have a higher survival rate in the breeding season, but not a higher average number of young fledged; the larger groups produce more fledglings although not in proportion to the number of females in the group. The eggs of early-laying females are tossed from the nest by the later-laying female, who may be socially dominant in the breeding group, in observations where eggs were identified by size and shape (rather than by molecular genetics, not yet determined in this species) and where social interactions between females were seen at the nest (Vehrencamp 1977,Vehrencamp et al. 1986). Males brood at night, females in the daytime.Although the males that mate with dominant females may leave more young (their mates remove the eggs and young of other females), these males are more likely to be taken by a predator. One risk is the nocturnal carnivorous bat Vampyrum spectrum which find anis on the nest and in roosting sites and kills and eats them (Vehrencamp et al. 1977). During the daytime, brooding females on the nest are occasionally fed by another ani, apparently her mate. Juvenile anis when reared in an earlier brood sometimes feed the young in the second brood (Skutch 1959, 1983).

After anis fledge and are independent, the young birds usually disappear from their group and move out of their natal area. Nevertheless a few males remain as nonbreeders or become breeders in their natal group, when another bird disappears or when their group size enlarges. Most anis do not breed in the group where they were born, and when they do it is only males that stay and breed in their natal group (Bowen et al. 1989, Koford et al. 1990). 3. Smooth-billed Anis breed in communal nests where as many as 10 adults attend to a nest and only one adult broods at a time. Eggs are sometimes layered in the nest with the lower layers covered with leaves (Euler 1867). Breeding females compete for the position of eggs in the nest.The females that lay later cover the eggs of the early-laying females with green leaves. Egg burial occurs in nests where two or more females lay. As a result, many eggs that are laid early do not hatch while eggs of the laterlaying females have a higher hatch success. The young from early eggs have a head start in development and are larger in a mixed-age brood, and young of the last-laying females are more likely to starve. In some nests, the last-laying female and her mate desert the nest, and their young are reared by the earlier-laying pairs (Loflin 1983). Females stay on the nest and are fed by their mates (Skutch 1983, Köster 1971). This cooperative behavior between members of a pair may allow the female to remain on the nest and guard her eggs from her rivals.The fledged young often remain in the group for a season and help care for later broods of their natal group as sentinels, and occasionally feed them. Groups with many anis tend to have larger territories, perhaps because the extra birds can exclude other anis from the group territory (Quinn and Startek-Foote 2000). In addition to communal nesting in their social group, Smooth-billed Anis sometimes lay in nests of other ani social groups.A female may lay in another group’s nest when her own nest is lost due to predation or to bad weather. These intruding females and their mates are attacked by resident breeding anis. If the intruders succeed in laying, their eggs are often buried by the resident group if they are laid early, but are incubated if they are laid when the last resident female has begun to lay (Loflin 1983).

Cooperative breeding 135 And anis sometimes nest as single pairs with one female laying in the nest, as they regularly do in Florida (Loflin 1983). All cooperatively breeding cuckoos not only nest together, they are social through the year, remaining in groups when they feed and roost. They perch pressed in contact against each other, and they roost piled in layers on the backs of their fellows.When they are in contact side by side, they preen each other (Sclater and Hudson 1889, Wetmore 1926, Barbour 1943, Wetmore 1968, Skutch 1983, Figure 9.1). Allopreening in cuckoos is known only in these social species, so it appears to be a social adaptation for living in groups.These birds have a high rate of infestation by parasitic insects, and allopreening perhaps is a social adaptation to the shared mallophaga or feather lice that are transmitted among members of the social group, much as allopreening is in Old World primates. Mallophaga are also common in their nests, and in Guira Cuckoos where a group may have as many as five nests in a season (R. Macedo, in litt.), most groups move to a different tree for the second nest.This change of nest site may decrease the nest infestation by the feather lice and other insects or nest detection by predators (Abrahamovich and Cicchino 1985). Cooperative breeding in these

cuckoos is of concern not only because of conflicts of interest between adult females in the group, but also because of the functional overlap of cooperative breeding and brood parasitism. In both the cooperative and brood-parasitic breeding styles, one bird cares for the young of another to which it is not closely related. One question in all these cooperative-breeding cuckoos is their relatedness, where birds in a social group may not be close kin. In Groove-billed Anis, birds often disperse from their natal area, and as a result the breeders may not be close relatives (Vehrencamp et al. 1986), and Guira Cuckoos in a group are not closely related, at least in a preliminary genetic study (Quinn et al. 1994). In Smoothbilled Anis, genetic markers provide evidence of complex relationships, with suspected cases of polygamy, extra-pair fertilization, intraspecific brood parasitism, or all these behaviors (Quinn and Startek-Foote 2000). Living in a group may benefit the individual birds in the group, and helping behavior may be a social mechanism that allows these birds to remain in the group and be tolerated by the other group members. A second question for future studies is the reproductive success of each female in a group. Early field studies estimated the contributions of each

Figure 9.1. Social behavior in group-living Guira Cuckoos Guira guira (after Fandiño Mariño 1986).

136 The Cuckoos female cuckoo by size and shape of the eggs (Vehrencamp 1976, Loflin 1983,Vehrencamp et al. 1986), and not by the more reliable molecular markers of identification of eggs laid and matched to individual females (Cariello et al. 2002). A third question is whether one female controls the reproductive success of other females in a group, or whether each female has the incentive to succeed within a changing social system (Reeve et al. 1998, Johnstone et al. 1999). In Guira Cuckoos the sequence in which females lay within a communal nest can change from nest to nest, and even the last-laying females sometimes lose eggs, so there is no obvious long-term social rank in relation to laying sequence (Macedo et al. 2004). On a broader scale, these cooperatively-breeding cuckoo species are each other’s closest relatives. Even though they live in different habitats ranging from the swamp forests of Greater Ani to the scrubbier habitat of Groove-billed Ani and Smooth-billed Ani

and the savannas of Guira Cuckoo, the crotophagine cuckoos share features of social behavior, nestbuilding, competition and cooperation among females, and infanticide. These cooperative cuckoos are an evolutionary parallel with the Australian fairy-wrens, a clade in which all 13 species are cooperative breeders and live in habitats that range from scrub to rainforest to desert. In fairy-wrens a sufficient phylogenetic explanation of their behavior is that their common ancestor was a cooperative breeder (Russell 1989, Edwards and Naeem 1993, Rowley and Russell 1997). This explanation is more likely than the idea that cooperative breeding is somehow caused by a common habitat across all species with this behavior, as there is no common ecological feature that is associated with the behavior. The same perspective applies to the crotophagine anis and Guira Cuckoo. In one sense, the most likely explanation of why each is a cooperative breeder is that their common ancestor had this behavior.

10 Brood parasitism

Brood parasitism is the most fascinating feature of cuckoo biology. Cuckoo parasites use other species to rear their young, and how the behavior evolved as a reproductive strategy demands an explanation (Darwin 1859, Hamilton 1972). Brood-parasitic cuckoos lay their eggs in the nests of other species, their hosts, and the nesting pair that rears the cuckoos are the foster parents.These birds incubate the cuckoo eggs along with their own, and when the cuckoo egg hatches they feed the young cuckoo. The foster parents forfeit their own reproductive success when the nestling cuckoo removes the hosts’ eggs or nestlings. Some cuckoo eggs mimic the background color and spot pattern of the eggs of their host, and the eggs of a single species of cuckoo may vary among females, each matching the eggs of a different species of host. Host egg mimicry in cuckoo eggs has been known for a long time. In Europe, the eggs of Common Cuckoo Cuculus canorus often match the eggs of the host nest where they lay, as noted in 1850 (Kunz 1850), described in detail by Baldamus (1853, 1892), and explained as a result of selection by the host, which removes from the nest any eggs unlike its own, by A. Newton (1896). In fact, Newton cited naturalists’ knowledge of cuckoo egg mimicry as early as the second century CE. Later biologists have reported new observations and developed ideas about the evolutionary process of natural selection in brood parasitism (Rey 1892, Swynnerton 1916, Rensch 1924, Makatsch 1937, 1955, Baker 1942, Brooke and Davies 1988, Davies et al. 1989a, Øien et al. 1995, Rothstein and Robertson 1998, Davies 2000). The Common Koel in India was the first bird described as a brood parasite. Four thousand year

old Sanskrit literature referred to these birds as “anya-vapa”; “anya” indicating “other” and “vapa” indicating “reared” or “brought up”; the term “anya-vapa” means the koels were “reared by others”. In a later Vedic drama, the Sakuntala by Abhijnanashakuntalam, dated about 375 AD, is the passage, “Koels manage to have all their young reared by other birds before they move into the air.” The same drama refers to the “other” as House Crow Corvus splendens, called “para-bhrit,” or “one who rears others” (Friedmann 1964b). Later the Moghul Emperor, the naturalist Jahangir (1569–1627) wrote: “A strange thing about the Koel is that it does not bring up its young from the egg, but finding the nest of the crow unguarded at the time of laying, it breaks the crow’s eggs with its beak and throws them out, and lays its own in the place of them and flies off. The crow thinking the eggs its own, hatches the young and brings them up. I have myself seen this strange affair at Allahabad”. Jahangir also described the Jacobin Cuckoo as a brood parasite: “I have seen in Kashmir that the papiha [the Jacobin] lays its eggs in the nest of the ghaughai [a babbler] and the ghaughai brings up its young” (Ali 1927). Brood parasitism in Common Cuckoos in Europe has been common knowledge since the classical period of the Greeks and was described in the natural histories of Aristotle. The evicting behavior of nestling cuckoos and other broodparasitic behaviors of these birds in Britain was described in detail more than 200 years ago ( Jenner 1788). In addition to these accurate observations, there have been misunderstandings concerning these cuckoos. “Cuckoldry” derives its meaning

138 The Cuckoos from a misunderstanding of cuckoo parasitism. Centuries ago it was thought that the female cuckoo mated with the male of the foster species, and this would “explain” both the color and the small size of the cuckoo’s eggs (Stresemann 1934, Rothschild and Clay 1952). It was common folk knowledge that the cuckoo “changeling” in the nest— substituted for the real offspring of the pair—is not the offspring of the pair that gives the parental care. In the contemporary sense of cuckoldry, rearing the young of another bird is what happens when a female mates with one male and her social partner rears a youngster that is genetically not his own. References to cuckoo changelings in Shakespeare’s plays were concerned with parents rearing young other than their own offspring, in an age when civil wars were fought and England was divided over kinship claims; the question of parentage and royal succession was a matter of life and death (Payne 1997b).This concept of cuckoldry has been central to the development of ideas in the late twentieth century about the evolution of parental care (Hamilton 1990). In the New World, brood parasitism in cuckoos was first reported about a hundred years ago (Friedmann 1964b), a century later than brood parasitism was described in the cowbirds. Hartert and Venturi (1909) mentioned an American Striped Cuckoo Tapera naevia parasitizing a spinetail Synallaxis nest in Argentina, but they did not imply that this was news to them, so brood parasitism may have been documented earlier. Azara (1809) mentioned nothing about parasitism of the New World cuckoos, and apparently the earliest record is from Suriname (then, Dutch Guiana) when the Penard brothers collected eggs of this cuckoo in the nests of other birds from 1898 to 1904 (Penard and Penard 1908). The next revelation of broodparasitic cuckoos in the New World was an observation by von Ihering (1914). He found a strange egg in the nest of a Black-backed Water-Tyrant Fluvicola albiventer in 1913, and guessed that it may be the as yet unknown egg of Pheasant Cuckoo Dromococcyx phasianellus.A female cuckoo taken the next year had in her oviduct an egg that matched the egg in the tyrant nest, clearing up the identity of these eggs (Naumburg 1930).

Identification of cuckoo eggs is a problem for field ornithologists as well as for the nesting hosts. Skead (1951) recognized the problem in South Africa,“When attributing cuckoo eggs to a particular species in an area where there are several species of cuckoos one is faced with the inevitable question, ‘How do I know they belong to that species?’ Unless laying is seen the only way to answer the question is to allow the egg and its contents to mature. But predatory Nature so often intervenes that one’s efforts are usually frustrated. I feel as sure as it is possible to be without absolute proof that the eggs of the Black-crested Cuckoo are pure white, shiny and nearly round. All those found in bulbul nests have been thus and have contrasted strikingly with the eggs of their hosts both in size and colour.” In India where several cuckoo species occur and two or more may overlap in their use of hosts, it is uncertain which eggs go with which species of cuckoo. The eggs identified by Baker (1906, 1907, 1908, 1934, 1942) and other egg collectors may not have been correctly identified. Even though Baker described eggs taken in the oviduct of females, the cuckoo itself was rarely saved as a museum skin, so doubts remain. Also, eggs that were found in host nests did not hatch, and the eggs collected were labeled years after they were taken from the nest (Harrison 1969, Becking 1981). A cuckoo egg can be identified by measuring and photographing the egg, then letting it hatch and develop into a feathered young cuckoo; or by sampling the eggshell or the nestling, with a feather or a drop of blood to compare its genetic sequence with that of a known cuckoo, as we have done with brood-parasitic finches (Payne et al. 2000, 2002). Identification of nestling and juvenile cuckoos has been a problem as well. Some errors, as in juveniles of Plaintive Cuckoo being mistaken as juveniles of Violet Cuckoo (Baker 1906, 1927,Wells 1999), can be traced back to earlier plumage descriptions, in this case to Hume (1875). Only recently have cuckoos been handreared from the egg or nestling stages to determine the juvenile plumage (Higuchi and Sato 1984). Fieldwork is needed to determine the basic life history details of many brood-parasitic cuckoos. The mysteries of brood parasitism in the cuckoos have led to evocative book titles such as

Brood parasitism 139 Cuckoo problems (Baker 1942), The Cuckoo’s secret (Chance 1922) and The truth about the Cuckoo (Chance 1940). We now phrase these mysteries as biological questions, and ask how birds without a family life of their own know what species they are? And, how they can find a nest where their own young will be reared? These questions lead us to consider how birds recognize their own young, and why the hosts rear young that are not their own. Do the brood parasites have a real effect on the breeding success of their hosts? Do the hosts distinguish their own young and the young of the cuckoo? Why do the hosts care for these foundlings? How have cuckoos evolved eggs that mimic the color and pattern of the eggshell of their host species, and how do their hosts respond to this challenge? Finally, how did the behavior of leaving their young in the nests of other birds evolve in the first place?

Behavior of the female cuckoo How females find a nest and lay in it During the breeding season female Common Cuckoos often perch on high branches and watch small songbirds as these make trip after trip with nest material to their nest sites. Nests of a host species are more likely to be parasitized when they are near regular cuckoo perches than when far from a cuckoo perch (Øien et al. 1996). Female cuckoos also search through the vegetation where songbirds are likely to nest. Cuckoos respond to calls of the hosts and may find a nest as the host gives alarm calls that increase in number and loudness when the cuckoo gets closer to the nest (Seppä 1969). We too find songbird nests in the same way by keeping track of alarm calls of the nesting birds (Payne and Payne 1998b). A female cuckoo visits the nest when the host is not there. She often takes an egg of the host in her bill, or she perches on the nest rim and swallows egg after egg as she takes them from the nest. She then sits on the nest and lays her egg directly into the nest. Egg-laying is rapid, 10 seconds or less; then the female flies away. With the best of timing she lays when a host egg has been laid but before

incubation has begun (Gurney 1897, Chance 1922, Molnár 1944, Gärtner 1981a,b, 1982, Wyllie 1981, Yoshino 1988, 1999). Edgar Chance first filmed Common Cuckoos in the nests of Meadow Pipits Anthus pratensis at Pound Green Commons in Worcestershire, England, in 1918–1921. By knowing where the pipits were nesting and by disrupting a few nests causing pipits to nest again at a time convenient for himself and the cuckoo, Chance was able to anticipate where the female cuckoo would visit and lay, in late afternoon on every other day, and he was there with his camera before the female visited the nest (Chance 1922). In earlier years it was thought that a female cuckoo laid her egg elsewhere and carried it to the nest in her bill, but reports of a cuckoo carrying an egg in the bill likely were of a female as she removed a host egg from the nest. In Britain, a committee was appointed to solve this cuckoo problem (Snow 1992).There was no final committee report, nor was one necessary: Chance’s films documented that cuckoos lay directly into the nest. Adapting current field methods in their observations, a Norwegian research group set up videocameras to make continuous recordings at nests of Reed Warbler Acrocephalus scirpaceus every day through the laying period.Videocameras were often in the right place at the right time: Common Cuckoos were recorded as they laid in the monitored nests (Moksnes et al. 2000). Cuckoo behavior here was similar to the behavior in England, although the female did not avoid visiting a nest when the host was nearby, and she often laid in the presence of the host. Cuckoos took host eggs from nests that were never parasitized as well as nests where the cuckoo laid. Parasitized warblers were more likely to remove the cuckoo egg from their nest when they had seen the cuckoo at their nest. In video recordings of Horsfield’s Bronzecuckoos Chrysococcyx basalis in Australia, a female visits the host nest alone and sometimes lays in a nest that she cannot enter. She glides from a perch to the nest site and lays while perched on the rim of the nest, or perched above it. In nests that are enclosed or in a hole the female pushes her abdomen into the nest cavity and lays directly into the nest (Brooker et al. 1988).

140 The Cuckoos In a few cuckoo species the female lays in the host nest with the help of her mate. In Jacobin Cuckoo Clamator jacobinus in South Africa the male often distracts the host from the nest, the host chases the male cuckoo, and the female slips into the unguarded nest and lays her egg (Liversidge 1971). In Common Koel Eudynamys scolopacea the male may distract the crow and allow the female to lay in the nest, though she usually visits the nest alone (Lamba 1969, 1976). Channel-billed Cuckoo Scythrops novaehollandiae pairs sometimes approach the nesting host together; the male distracts the host while the female lays in the unattended nest (Crouther 1980, Larkins 1994b).

Why cuckoos parasitize certain species Several biological and ecological factors affect the suitability of bird species as hosts (Sealy et al. 2002). These include whether a potential host species is likely to accept the cuckoo egg, incubate it, care for the young and meet its nutritional needs. Insectivorous songbirds with short nestling periods provide the best parental care because these hosts bring food for the nestlings at a high rate. In addition, the number of host pairs in the local population may affect the choice of a host, as a common host provides more nests for a cuckoo to parasitize. In a comparative study of songbird species available to Common Cuckoos in Europe, Soler et al. (1999b) determined that nests of the more abundant species were more likely to be parasitized. Other factors that affect the rate of parasitism were the length of nestling period, the time of breeding season, and the nest structure and location (open nesting species are more likely to be parasitized than cavity nesting species). Friedmann (1967) suggested that competition among cuckoo species for parental care may have led the brood-parasitic cuckoos to focus on different hosts. Although many cuckoos are host specialists, in some regions two or more cuckoo species overlap in their choice of hosts. A nest of Western Thornbill that we found on Gooseberry Hill in Western Australia had two cuckoo eggs, one an unspotted copper egg of Shining Bronze-cuckoo

and the other a spotted whitish egg of Horsfield’s Bronze-cuckoo (Payne and Payne 1998a). Splendid Fairy-wrens are used by Shining Bronze-cuckoo and Horsfield’s Bronze-cuckoo in South Australia; Jungle Babblers Turdoides striatus are used by Jacobin Cuckoo and Common Hawk-cuckoo in India (sometimes both kinds of eggs appear in the same nest: Baker 1906); some warblers and sunbirds are used by two glossy cuckoos Chrysococcyx species in Africa, and so on (Baker 1942, Payne and Payne 1967, Becking 1981, Brooker and Brooker 1989a, Brooker 1992).There is no evidence that a cuckoo species uses more host species in the absence of another cuckoo species than when it is the only brood-parasitic species, as would be expected from the competition hypothesis. Also two species of brood-parasitic cuckoos do not respond to each other’s songs as would be expected if they were strongly competitive. The host specificity of cuckoos is better explained by the success in parasitism of each host species by a cuckoo species, survival of cuckoo eggs and nestlings, imprinting, and in some species the coevolutionary match of cuckoo eggs and host eggs, than by interactions between cuckoo species.

Does a female imprint on her foster species? Imprinting is a behavioral process in which the young bird forms a social attachment to an object to which it directs its adult behavior (chooses a mate, or seeks and parasitizes a nest of a socially familiar kind of bird), an attachment that occurs during a critical period in early life, is irreversible, and is generalized to other individuals of the species (Lorenz 1935, Irwin and Price 1999, ten Cate and Vos 1999, Payne et al. 2000). Maintenance of the egg morphs of females in the egg-mimetic cuckoos may involve imprinting. Female Common Cuckoos in Britain and Europe lay eggs that differ from the eggs of other females, and often each female’s eggs match the eggs of a host songbird (Baldamus 1892, Rey 1892). Each female normally lays in the nests of a single kind of host (Chance 1922, 1940, Marchetti et al. 1998). How does a female decide which host to parasitize?

Brood parasitism 141 Imprinting to its own foster parent’s species might be the behavior that maintains a tradition of host specificity in brood parasitism from one generation to the next. A young cuckoo could imprint to her foster species, so that her adult behavior in host choice would be directed towards the species that had reared her. If she formed an association with this host species, the cuckoo may return as an adult to the same habitat and find nests by recognizing the host species and matching a memory of calls of her own foster parents to the calls of nesting birds, then lay in those nests. According to the “gens” theory, a population of Common Cuckoos consists of distinct biological “races” or “gentes” (Newton 1896) that are closely associated with their foster species and mimic the egg color and pattern of this host species. The female cuckoos then seek and choose the nest of this species to lay their eggs. A female cuckoo is host specific, by this reasoning, because she was behaviorally imprinted to her foster parents. Her own experience as a chick directs a female cuckoo to search out the species that reared her and parasitize that same host species. By using a host like the one that reared her, the female chooses a host species that has reared a cuckoo, and she can lay an egg that matched the egg from which she hatched and that has a high probability of acceptance by the same host species. There has been little research with cuckoos to test the idea of imprinting. Löhrl (1979) kept wildreared Common Cuckoos in aviaries. A female reared by Pied (White) Wagtail Motacilla alba saw wagtails outside the aviary, and although she could not see the wagtail nest, the cuckoo laid an egg in the aviary two days before the wagtail laid. Other studies have tested cuckoos reared in different conditions, where the cuckoos had alternate foster species or had alternate habitats. First, Brooke and Davies (1991) cross-fostered nestling cuckoos from Reed Warbler nests into nests of alternate foster-parents, including European Robin Erithacus rubecula, which reared the young cuckoo to independence. Other young were reared by moving the nest, cuckoo and adult warblers into an aviary where the warblers reared the cuckoo. Next year the captive cuckoos were tested to see if they

perched near their own foster species. When they had a choice of Dunnock Prunella modularis, robin or warbler, the cuckoos showed no preference for perching near their foster species, even when one, a female cuckoo, was implanted with estradiol. Second,Teuschl et al. (1998) removed young cuckoos from nests of Reed Warbler and Marsh Warbler A. palustris, hand-reared them with different visual effects (“habitats”) in small cages, then kept the cuckoos in an aviary. Over the next two years the cuckoos were tested in a cage where each could perch near the six visual “habitats”. One cuckoo perched near its natal “habitat”, but in a second test she did not. Neither study tested whether a cuckoo imprints to her foster species, then lays her eggs in the nests of the same host species. The imprinting model is a reasonable explanation for the host-specific preference of female cuckoos. However, the difficulty of breeding cuckoos in captivity, and the dispersal of cuckoos from their natal site to first breeding site in the field, have prevented a convincing test of the imprinting model. The imprinting model has been tested in another broodparasitic bird, Village Indigobird Vidua chalybeata. The indigobirds are host-specific brood parasites, and their nestlings mimic the nestling colors of their normal host species. Indigobirds were foster-reared in aviaries, either by their normal host species, the Redbilled Firefinch Lagonosticta senegala, or by an experimental foster species, the Bengalese Finch Lonchura striata. Captive-reared female indigobirds were tested as adults for host choice. In the aviaries where both finch species were nesting, the females reared by firefinches laid in the nests of firefinches, and the females reared by Bengalese Finches laid in the nests of Bengalese Finches. Females caught in the wild showed the same behaviors as captive-bred females reared by firefinches.The results showed imprinting in which young indigobirds focus their attention on their foster parents, rather than an innate bias for the normal host species, or a nonspecific choice of a host. In these imprinting experiments the females recognized and laid in the nest of the species that reared them, even when the foster species was not the normal host species (Payne et al. 2000). Behavioral conservatism may explain why certain host species are parasitized in one region and not in

142 The Cuckoos others. But even in a population where all females imprint to their normal host species, a female may occasionally lay in the nest of a novel host, as she does when a nest she has been observing is destroyed before the cuckoo has laid; she may then lay in the nearest available nest of another species, which may rear the young cuckoo (Rose 1982). In southern Germany where Common Cuckoos parasitize other host species, 3000 nests of Blackcap Sylvia atricapilla were checked before one was found with a cuckoo; the Blackcap pair reared the young cuckoo (Berthold et al. 1995). In another species, Great Spotted Cuckoo in Spain parasitize Magpie Pica pica and Carrion Crow Corvus corone.The cuckoo eggs in the nests of these two hosts look alike. Some female cuckoos lay eggs (identified by genetic markers that matched the adult females) in nests of both species of host, Magpies as the common host, and Carrion Crows as alternative host when no Magpie nests are available (Martínez et al. 1998a). The same cuckoo species has colonized new host species when these species have been introduced into the normal range of the cuckoos, the House Crow in Israel and the Indian Mynah in South Africa.

Egg color and pattern, and the genetics of mimicry Common Cuckoo eggs often resemble the color and pattern of their hosts’ eggs.The eggs of this cuckoo vary in color and pattern within a population. Some eggs are similar to eggs of one host species and others are similar to those of another host species (Baldamus 1853, 1892, Rey 1892, Oates 1903, Makatsch 1937, 1971, Yoshino 1999). Many hosts recognize the eggs of a cuckoo as different from their own, and these hosts remove or “reject” the cuckoo egg from their nest if it does not resemble their own eggs. The function of cuckoo egg matching is thought to increase the change of parental care by the nesting host. Cuckoo eggs that mimic the color and pattern of the host eggs are more likely to be accepted by the host and incubated along with the host’s own eggs. In field experiments, the mimetic eggs are less likely to be rejected from the nest than are eggs that differ from the host eggs (Swynnerton 1916, 1918, Davies and Brooke 1988, Davies 2000).

Fine tuning of cuckoo egg to host egg by natural selection may lead to specialization, with eggs that match one species of host.There may be a coevolutionary “arms race”, where the host is selected for an ability to discriminate among eggs, and the cuckoo is selected to lay an egg that mimics the hosts’ eggs (Baker 1907, 1942, Jourdain 1925, Livesey 1936, Southern 1954, Davies 2000). In Britain many female Common Cuckoos lay distinct eggs that closely resemble the eggs of the most common host species (Davies and Brooke 1989). Mimicry of the cuckoo eggs to the host eggs is not perfect, and the cuckoo eggs are still distinguishable as cuckoo eggs by color and pattern as well as by size and shape (Cramp 1985: plate 96). In Central Europe the cuckoo eggs laid in the nests of four host species of Acrocephalus warblers all look alike and do not match the differences between the eggs of these host species (Edvardsen et al. 2001), and in Japan the same is true for cuckoo eggs laid in the nests of different host species (Nakamura et al. 1998). In surveys of museum egg collections in Europe only 25–50% of cuckoo eggs are a close match to the host eggs (Harrison 1968, Perrin de Brichambaut 1993). Several authors have noted a difficulty in distinguishing cuckoo eggs from host eggs, and the low match in museum collections may result in part from a bias of collectors having passed over the clutches they did not realize had a cuckoo egg when the egg was a good mimic. Although eggs of Common Cuckoos do not always mimic the color and pattern of the host eggs, the eggs that match have a greater chance of being accepted by the nesting host (Davies 2000). When models of cuckoo eggs were tested in the host nests, certain hosts were more likely to accept the model egg if it was like the cuckoo egg normal for their nest. In nesting hosts that are parasitized by mimetic cuckoo eggs (Meadow Pipit, Pied Wagtail, Reed Warbler), the model eggs were either accepted or rejected (the eggs were removed from the nest, or the nest was deserted). Meadow Pipit and Reed Warbler accepted the model eggs that matched their own cuckoo egg morph, but rejected the other eggs. Pied Wagtail rejected the model eggs half the time when these matched their own eggs and more than half when they did not match, but

Brood parasitism 143 the difference was not significant.The fourth host, European Robin, is parasitized by cuckoo eggs, like the cuckoo eggs in nests of Pied Wagtails (much like the robin eggs); while the fifth host, Dunnock, is parasitized by spotted whitish cuckoo eggs, although Dunnock eggs are unspotted blue. These last two host species accepted model eggs whether or not they were unlike their own eggs (Davies and Brooke 1989a).The tests demonstrated the importance of egg mimicry in acceptance of cuckoo eggs by certain species of nesting hosts. The tests also demonstrated the discriminating ability of the hosts. Avoidance of predation has been proposed to explain the function of egg mimicry.Wallace (1889) suggested that cuckoo eggs in the nests of birds with matching eggs would be less conspicuous to a predator. Egg predation could involve other cuckoos; as when a female cuckoo lays in a nest, a second female would be less likely to remove the rival first cuckoo’s egg if the first cuckoo egg matched the host eggs (Brooke and Davies 1988, Brooker et al. 1989b, Brooker et al. 1990).There is no evidence of a higher rate of predation in clutches where a cuckoo egg was unlike the hosts’ own eggs. In field experiments there was a tendency for second cuckoos to remove a model cuckoo egg when it did not match the host eggs, but the tendency was not strong enough to be statistically significant. In addition, few nests have second cuckoo eggs, so the risk of egg predation by a second cuckoo is low and selection to match to escape egg predation by a second cuckoo would not be as strong as discrimination by the host (Davies and Brooke 1988). Female and male genetic contribution to the egg color and pattern of cuckoo eggs once puzzled ornithologists. Current genetic research has demystified this puzzle. A female bird produces the color and pattern on the egg while the eggshell is in her oviduct; the male that fertilizes the egg has no effect on the appearance of the eggshell (Brooke 1989). Certain genes are transmitted to the offspring only by females and not by males. One set is on the sex chromosomes, a pair that differ in size and shape in one sex but match in the other sex. In most birds, including the cuckoos, females are heteromorphic for sex chromosomes (females are WZ, males are

ZZ), in contrast to most mammals in which males are the heteromorphic sex (males XY, females XX) (Punnett 1933, Ray-Chaudhuri 1967, 1973, Waldriguez and Ferrari 1979, 1980, 1982,Waldriguez et al. 1983, García-Moreno and Mindell 2000, Graves and Shetty 2001). In birds, genes that are transmitted from mother to offspring (and not from father to offspring) occur on the W chromosome (Graves and Shetty 2001). Genes that are transmitted in this way also occur on mitochondria, outside the nucleus (Avise 1994). Genes that are transmitted only along female lineages behave much like the female brood parasites do when they lay their eggs in nests of the same kind of hosts from one cuckoo generation to the next (Sorenson and Payne 2002). In mammals the X chromosome carries about 10% of the nuclear genome and has an estimated 3000 genes; the short Y has fewer, some with counterparts on the X and some unique to the Y and expressed only in males (Graves and Shetty 2001, Klein and Takahata 2002). In birds, the Z comprises 7% of the nuclear genome; the W has much less. In the only bird (chicken) that has been gene mapped, the DNA of the terminal PAR (pseudoautosomal) region of the W and Z are the only parts that recombine. Genes on W and Z differ in sequence detail and do not recombine in meiosis (Fridolffson et al. 1998, Ellegren 2000, Graves and Shetty 2001), and so the genes on the nonrecombining region of W are transmitted only from mother to daughter. In principle, genes coding for traits that are expressed only in one sex should disproportionately have their addresses on these sex chromosomes (Roldan and Gomendio 1999). It is unknown whether W genes affect in birds the color end pattern of eggshell pigments.The mitochondrial genes are unlikely to code differences in eggshell color and pattern. The coding functions of mitochondrial genes are well known: mitochondrial gene content consists of 2 ribosomal RNA genes, 22 transfer RNA genes, and 13 protein genes that code for subunits of enzymes that function in electron transport and ATP synthesis (Moritz et al. 1987, Desjardins and Morais 1990, Mindell et al. 1999). In birds the eggshell pigments are porphyrins rather than melanins, and apart from domestic fowl there is no information about the

144 The Cuckoos developmental genetics of these porphyrin pigments (Lang and Wells 1987, Miksik et al. 1996). Do mother and daughter birds lay the same kind of egg? There are no observations in cuckoos (Hamilton 1972), but there are in another bird. Great Tit Parus major daughters have eggshell patterns like mother and maternal grandmother, but not like paternal grandmother (Gosler et al. 2000). The observations are consistent with the idea that eggshell pattern is inherited along a maternal lineage in these birds. To test whether Common Cuckoos reared by different hosts differ genetically, and whether cuckoos reared by the same host share the same genetic profile, Gibbs et al. (2000) compared mitochondrial genes of cuckoos in the nests of their main host species. In Britain, nestling cuckoos were sampled in the nests of Reed Warbler, Meadow Pipit and Dunnock. Each variant gene was found mainly in nestling cuckoos with a single species of host. These genes differed between localities, and cuckoos in the nests of one host species (Reed Warbler) in different localities shared some gene variants. Because mitochondrial genes are inherited along the maternal line, the variation among cuckoos in the nests of a host species indicates that female cuckoos have switched hosts from time to time. It also indicates that some races of the cuckoos have more than one origin, perhaps from times when separate cuckoo lineages colonized the same host species: multiple colonizations can account for the different genes in cuckoos in the nests of the same host species. In Japan, nestling cuckoos were sampled in the nests of Great Reed Warbler and Azure-winged Magpie Cyanopica cyanea. Cuckoos in magpie nests shared the same gene variant with each other. Cuckoos in warbler nests had four variants including the one in cuckoos in magpie nests. Warblers are parasitized through much of their range, whereas magpies are parasitized locally: the genes suggest that magpie-cuckoos were descendants of a single female. In contrast to the mitochondrial genes, the nuclear microsatellites did not show significant differences between cuckoos in nests of different host species in Britain or in Japan (Gibbs et al. 2000). The genetic results are consis-

tent with the field observations that female Common Cuckoos focus on a particular host species and males do not, and females do not tend to mate with males with the same genetic variants (Marchetti et al. 1998). A few other species of brood-parasitic cuckoos have variable colors and markings on their eggs (Plate 17). The Australian cuckoos including Horsfield’s Bronze-cuckoo, Shining Bronze-cuckoo and Black-eared Cuckoo have eggs that are similar to some of their host species, but only the Brush Cuckoo has polymorphic eggs. In southern Asia and Africa, the Jacobin Cuckoo eggs are of a similar blue color to eggs of their host species, but in southern Africa the eggs are unmarked white, unlike any of their host species. Egg color and pattern in the Black Cuckoo appears to be similar across its range in Africa, and these birds parasitize a species group of hosts: bush-shrikes Laniarius that all have similar eggs, usually bluish green with indistinct purplish or brownish spots and blotches across the different species (Fry et al. 2000, Plate 18d). Red-chested Cuckoo eggs vary geographically, unmarked dark brown in southern and East Africa where they parasitize robin-chats Cossypha with brown eggs, and pale blue, unmarked or with fine spots, in other parts of East Africa where they parasitize scrubrobins Cercotrichas; they do not always mimic the scrub-robin eggs (Plate 18e,f). The Chrysococcyx glossy cuckoos have polymorphic eggs, best known in Diederik Cuckoos with eggs of several colors and patterns. Female diederiks lay in the nests of a host with an egg color and pattern like their own (Hunter 1961, Payne 1967, Reed 1968, Jensen and Vernon 1970). Red Bishops Euplectes orix which lay unspotted blue eggs reject model cuckoo eggs that have dense spots or a darker background color than their own eggs (Lawes and Kirkman 1996).Village Weavers Ploceus cucullatus are host species that lay a variety of egg colors and patterns in a single colony (Plate 19). The weavers discriminate against eggs that differ from their own eggs, removing them from the nest, and the proportion of rejections depended on the dissimilarity in color and pattern of the other eggs to their own eggs (Lahti and Lahti 2002).

Brood parasitism 145

How does a cuckoo know its own species? The young brood-parasitic cuckoos do not grow up with their own parents, and they have no early opportunity to learn their own species from other cuckoos.When they are adults, the males apparently develop their song by themselves without having to hear it from an adult. Female mate choice behavior may develop normally whether or not they are exposed to that of others of their own species. Male cuckoos do not imitate the songs or other behaviors of their foster parents, and in consequence the female has no information of whether males were reared by the same host species that reared her.As far as is known, a female innately recognizes a conspecific male as a potential mate. Then on the basis of his song rate and other behaviors that indicate his physical condition and ability to defend an area, and on his attraction to the female and the interaction between the pair, she decides to mate with him. Although they are not reared by their own family, the social behavior of certain brood-parasitic cuckoos might be affected by other cuckoos, either their nestmates or older cuckoos. Great Spotted Cuckoos often lay more than one egg in a host nest (Friedmann 1964) and the young cuckoos are often reared together in the nest. Soler and Soler (1999) placed young cuckoos into the nests of Magpies in an area where the young had no social contact with adult cuckoos.When it was reared alone in a nest, a fledgling did not join other fledged young cuckoos; but when it was reared together with another cuckoo, it associated with other fledged young cuckoos. An influence of early social life on the mate choice of adult cuckoos is unknown. Because Great Spotted Cuckoos are sometimes reared without another cuckoo in the nest, these cuckoos can select a conspecific mate whether or not they experience early life with another cuckoo. In most species of brood-parasitic cuckoos the young grows up alone in the nest, and it evicts any eggs and young of another cuckoo or the host, or the nestling cuckoo kills the host young with the hooks on its bill. As a consequence, we think that adult cuckoos mate with conspecific adults

whether or not they interact with other cuckoos when they are young.That is, the choice of a conspecific mate appears to be innate in the cuckoos.

Behavior of the nestling cuckoo Eviction behavior The nestling in many species of brood-parasitic cuckoos actively removes the host eggs from the nest. After a female cuckoo removes a host egg from the nest, the infanticidal nestling cuckoo evicts the remaining contents of the nest.The cuckoo evicts the host eggs within a few hours of hatching, and if it fails then it succeeds in a couple days when it is larger and strong enough to lift and push the eggs from the nest.The cuckoo braces each foot against the bottom of the nest and braces its head by pushing the bill into the nest, forming a stable tripod. It balances the egg on its back and raises the hooked wings to each side to hold the egg in place. This behavior is aided by the flat or hollow shape of the nestling cuckoo’s back, and by the enlarged first digit which holds the egg in place on the nestling’s back.Then the nestling moves backwards and thrusts its rump upward and back until the egg falls free of the nest (Payne and Payne 1998a, Figure 10.1).

Figure 10.1. Eviction behavior of nestling Horsfield’s Bronze-cuckoo in a nest of Splendid Fairy-wren, Gooseberry Hill,WA.

146 The Cuckoos Fatal competition with host nestmates is won by the nestling cuckoo when it hatches first and the nesting parents do not interfere with the mayhem in their nest. When the cuckoo nestling removes the hosts’ eggs, the foster parents do not rear their own young.They give all their parental care to the cuckoo and rear it as their own. The cuckoo takes all the food that would have been delivered to the host brood. Eviction by the nestling is known in about half the species of brood-parasitic cuckoos. The behavior has been watched in most detail in the Common Cuckoo (Wyllie 1981, Khayutin et al. 1982, Malchevsky 1987, Yoshino 1999, Davies 2000) and it has been seen in several species of Cuculus, Cacomantis and Chrysococcyx and in Thickbilled Cuckoo Pachycoccyx audeberti. Evicting behavior is sensitive to the temperature in the nest. Unlike other altricial birds, the young cuckoos become more active at cool temperatures. Nestling cuckoos can evict at cool body temperatures, conditions when the host nestlings are motionless in the nest. In northern Russia, nestling Common Cuckoos are active and evict the host eggs when the nestlings are cool. “Temperature changes greatly affect the activity of the Cuckoo, which is most active and vigorous at the relatively low temperature of 15°[C], when the young of the host may be rigid with cold” (Dement’ev and Gladkov 1966). Promptov and Loukina (1940) found that the nestling cuckoo evicts better at lower temperatures, when its skin is cold.They suggested that this behavior is adaptive: the cuckoo evicts the host young when the host female is absent, the nest is cold and host chicks are inactive. Nestling cuckoos, like passerines, are inactive below 15°C, and are otherwise most active in begging when TB is 35–37°C, their temperature when they are brooded (Malchevsky 1987). Nestling Common Cuckoos begin to develop a high internal temperature about the age when they no longer evict, with temperature regulation beginning 4–5 days after hatching (Hund and Prinzinger 1980). In temperate Western Australia Chrysococcyx nestling cuckoos evict at low temperatures. Nestling Horsfield’s Bronze-cuckoo and Shining Bronze-cuckoo are active and evict at cool envi-

ronmental TA and body TB temperatures over a range of TB from 34°C to 20°C. Nestlings evict persistently when their TB drops 2–3°C and the nestlings are touched on the back, and they maintain this eviction push-up posture while the TB is above 20°C. When they are no longer touched on the back, they move around the nest with increased “searching” activity as they cool, and they resume the eviction posture when touched again on the back. Small nestlings (1.9 to 3.0 g) cool rapidly, TB dropping from 5.7° to 11.7°C below the initial TB of 30–32°C in a few minutes. Nestling cuckoos stop the eviction behavior when the TB drops to 20°C or cooler. They do not shiver, and they cool as rapidly as a host nestling. Nestling fairy-wrens of the same age (0–5 days) are cool at TA between 22° and 30°C and are motionless until they are warmed to 30°C (Figure 10.2).The nestling cuckoo remains inactive while it is brooded and kept warm, and when its foster parent leaves the nest for its first feed early morning, the cuckoo becomes active and pushes itself around the nest until it comes in contact with the eggs in the nest (Payne and Payne 1998a).This remarkable active behavior of the altricial parasitic cuckoo nestlings at cool temperatures that cannot maintain their own body temperature has no known counterpart in other altricial birds. New World brood-parasitic cuckoos also eliminate the competition for parental care. Nestling

Figure 10.2. Movements of young Horsfield’s Bronzecuckoo Chrysococcyx basalis in a nest of Splendid Fairywren Malurus splendens, Gooseberry Hill, WA. The nesting fairy-wren is immobile.

Brood parasitism 147 nestling cuckoo does not evict. In Spain, young Great Spotted Cuckoos in the nests of magpies grow at the expense of the host young, whereas young cuckoos in the nests of the larger crows often grow up together with their foster nestmates which do not have a lower success in these parasitized nests.

Begging behavior and begging calls

Figure 10.3. Hooks on bill of a five-day-old nestling American Striped Cuckoo Tapera naevia (after Morton and Farabaugh 1979).

American Striped Cuckoos Tapera naevia are siblicidal.They have a sharp hook on the bill tip and use the hook to kill their nestmates (Figure 10.3, Morton and Farabaugh 1979) much like a broodparasitic nestling honeyguide uses its bill hook to kill the nestlings of its foster parents (Friedmann 1955, Payne 1998).The other two brood-parasitic cuckoos in the New World, Pheasant Cuckoo and Pavonine Cuckoo, are less well known, though insofar as they are reared in an ovenbird nest that is more enclosed than the nest of the Australian fairy-wrens (a closed nest from which a young bronze-cuckoo can evict), the eviction of the egg or nestmates by these nestling neomorphine cuckoos would be difficult. Nestlings of other brood-parasitic cuckoos grow up together with their host young. In part, their tolerance of nestmates may be forced by the larger size of the host.The host-tolerant cuckoos include crested cuckoos Clamator spp. ( Jacobin Cuckoo, Levaillant’s Cuckoo, Great Spotted Cuckoo, Chestnutwinged Cuckoo), Common Koel (in India, where they parasitize crows), Channel-billed Cuckoo and Long-tailed Cuckoo. These tolerant cuckoos are sometimes reared alone, however, as they grow rapidly and out-compete the host young for food. In Australia where there is only one cuckoo egg in a nest and the host is smaller than the cuckoo, young Common Koels sometimes evict the host eggs. In India the young are often reared with another young Koel in the nest, the host is larger and the

The young parasitic cuckoo crouches in the nest, tosses the head back and opens the mouth. It calls loudly when the foster parents approach the nest, and calls more loudly and rapidly when a parent places food into its open mouth (Payne and Payne 1998a). It continues to beg with open gape after it is fed. The brood-parasitic nestlings differ from the nesting cuckoos in the mouth, which is unmarked in the parasites but often has species- or genus-specific colored patches in the nesting species (Plate 20). Nestling brood-parasitic cuckoos deposit their excreta in a gelatinous wrap as in passerine birds and the package of feces is carried away by the foster parents. After they fledge, the young cuckoo actively associates with its foster parents, flying towards them, calling, and begging with fluffed plumage, slightly drooped wings and open mouth (Figure 10.4). The nestling behavior is similar to the behavior of young songbirds and is remarkably different from that of the young of nesting cuckoos. In Yellow-billed and Black-billed Cuckoos, the nestling stands upright in the nest, stretches its neck upward and flaps its wings as early as its first day after hatching, and the young excrete a foul-smelling liquid (Hughes 1999a, UMMZ photos, Figure. 5.7). Young brood-parasitic cuckoos are loud and persistent in begging for parental care, both in species that are reared together with the host young and in species where the young cuckoo kills its nestmates and is the only nestling in the brood.The calls are so loud and given so often that the begging cuckoo makes as much noise as an entire brood of its host species (Davies et al. 1998, Davies 2000). Young cuckoos call when the foster parents are out of sight and also when they are nearby. Unattended fledged cuckoos use short calls to maintain social contact with the foraging foster parent. When the foster

148 The Cuckoos

Figure 10.4. Begging behavior of young Common Cuckoo Cuculus canorus, fed by foster parent Vinousthroated Parrotbill Paradoxornis webbianus: a, in nest; b, after fledging (after Yoon 2000).

parent approaches, the young cuckoo increases the rate and duration of its calls (Payne and Payne 1998a).The young of other brood-parasitic birds, the cowbirds Icteridae and brood-parasitic finches Viduidae, are also conspicuous in loud and persistent begging calls that not only attract their foster parents but also put them at risk of attracting a predator. Begging calls of most cuckoos differ from begging calls of their hosts, and the young cuckoos are readily identified in the field as cuckoos by their calls. Species-specific differences in mimicry of begging calls are unknown in the parasitic cowbirds (Gochfeld 1979, Woodward 1983, Broughton et al. 1987). Although some young brood-parasitic Vidua finches match the begging calls of their host, others do not (Payne and Payne 2002), and matching their

hosts’ begging calls is not common in nestling brood-parasitic birds. In a few cuckoos, the young give different begging calls when reared by different host species. A handreared young American Striped Cuckoo had a call like the begging call of the young of its host species, and it differed from the call of a cuckoo in the nest of another host species (Haverschmidt 1961, Morton and Farabaugh 1979). Fledged Levaillant’s Cuckoos are said to give calls like the begging calls of their babbler hosts ( Jubb 1952, Mundy 1973, Steyn 1973, Vernon 1982, Hustler 1997b, Barry 1998). Great Spotted Cuckoo young in nests of crows Corvus spp. and in nests of Magpies in Spain have different begging calls (Arias de Reyna and Hidalgo 1982, Redondo and Arias de Reyna 1988, Redondo 1993), and so do cuckoo young in nests of crows and nests of starlings in Africa (Mundy and Cook 1974). Begging calls of young Channel-billed Cuckoos in Australia are said to be similar to begging calls of their hosts. It is unknown whether female cuckoos are species-specific in host selection and have call races, or the nestling cuckoos learn their calls. Begging calls of Shining Bronze-cuckoos in New Zealand reared by Grey Gerygone Gerygone igata are identical to those of cuckoos in Western Australia reared by Yellow-rumped Thornbill and Western Thornbill (Payne and Payne 1998a). The host species differ from each other in their begging calls, so there is no evidence in this cuckoo for mimicry of begging calls as suggested elsewhere (McLean and Waas 1987). In another glossy cuckoo, Diederik Cuckoo in Africa, the young in nests of hosts Southern Masked Weaver Ploceus velatus, Red Bishop Euplectes orix and Cape Sparrow Passer melanurus are said to differ in begging calls (Reed 1968), and begging-call races may coincide with mimetic egg races in this cuckoo; recordings are needed to compare these begging calls. In Horsfield’s Bronze-cuckoo, the begging calls of the young differ when the nestlings are reared by different host species; and further, the begging calls of each are similar to the begging calls of the host young. On a hillside in Western Australia the begging calls of young cuckoos reared by Splendid Fairy-wrens at first are a “peep”, then change to “zeba” and “sit” calls and a “twe” note by late nestling

Brood parasitism 149 life; after fledging the young cuckoos combine “twe” calls into a “reel” like a young fairy-wren. Begging calls of cuckoos reared by Western Thornbills are a “whine” like the calls of a young thornbill.The begging calls of young reared by an occasional host, the irruptive Scarlet Robins Petroica multicolor which appear after grass fires, are like the calls of young cuckoos reared by thornbills and are not like the begging calls of young robins (Payne and Payne 1998a). Nestling cuckoos in the nests of fairy-wrens and thornbills had distinctive calls by the time they were 10 days old, and by 14 days the calls were the same as in the fledgling cuckoos.The rising buzz of young thornbills was not matched exactly by the cuckoos, nor was the descending buzzy note of robins, and some calls (“ze-ba,” “sit”) of cuckoos in the care of fairy-wrens did not match the calls of the host’s own young. Nevertheless, the begging calls of cuckoo and host species were similar, especially in cuckoos reared by fairy-wrens. Where the host-specific begging calls have been recorded in southeastern Australia (Buckingham and Jackson 1990), young Horsfield’s Bronze-cuckoos reared by Superb Fairy-wren Malurus cyaneus sound like young fairy-wrens and like the cuckoos reared by Splendid Fairy-wrens in WA, and young cuckoos reared by Brown Thornbill Acanthiza pusilla sound like young Western Thornbills and like the

cuckoos reared by thornbills in WA (Figure 10.5). Nestling Horsfield’s Bronze-cuckoos in nests of Superb Fairy-wren have a begging call like the missing young of the host as early as day 2 after hatching. The begging calls are thought to attract the parental care of the foster parents. In experiments, lone chicks of Superb Fairy-wren and of Horsfield’s Bronzecuckoo are less likely to be deserted than lone chicks of Shining Bronze-cuckoo, which have a different begging call (Langmore et al. 2003). Because nestling bronze-cuckoos evict the host eggs from the nest by day 2 after hatching, the young cuckoos have no opportunity to learn the begging calls of their host species from the foster parents’ nestlings. A possible explanation of the begging calls is that Horsfield’s Bronze-cuckoos have begging-call races, where the young of females that lay in the nests of fairy-wrens differ genetically from the young of females that parasitize thornbills. The irruptive host robins rear cuckoo young with begging calls like the calls of thornbill-reared cuckoos.The observation is consistent with the idea of a begging-call race of cuckoos that regularly use resident fairywrens and that opportunistically use other host species (Payne and Payne 1998a). It is not consistent with the idea that young cuckoos learn and copy the calls of their foster species, if only because the robinreared cuckoos did not call like the young robins,

Figure 10.5. Begging calls of young Horsfield’s Bronze-cuckoos Chrysococcyx basalis: a–c,Western Australia (a, cuckoo reared by Splendid Fairy-wren Malurus splendens; b, cuckoo reared by Western Thornbill Acanthiza inornata; c, cuckoo reared by Scarlet Robin Petroica multicolor); d–e, eastern Australia (d, cuckoo reared by Superb Fairy-wren M. cyaneus, NSW; e, cuckoo reared by Brown Thornbill A. pusilla,Victoria (d, e from Buckingham and Jackson 1991)); f–h, young hosts in Western Australia (f, Splendid Fairy-wren; g,Western Thornbill; h, Scarlet Robin).

150 The Cuckoos and because nestling cuckoos never hear the host young, as the cuckoo chick evicts the host eggs before they hatch. Bronze-cuckoos in some parts of Australia have mitochondrial genetic differences between populations, and so do cuckoos within a local population (Joseph et al. 2002). It will be of interest to test whether genetic differences are found between nestling cuckoos that are reared by different host species, and to transfer cuckoo eggs between the nests of different hosts and record the begging behavior of young cuckoos reared by these hosts. The matched begging calls of cuckoo chicks to calls of the host may occur only in certain glossyand bronze- cuckoos Chrysococcyx, insofar as Common Cuckoos in Europe do not differ in their begging calls when their host species differ (Butchart et al. 2003).

Ecology of host-parasite associations Unlike the young of nesting birds, the young brood parasites are not related to the adults that feed them. A young cuckoo can gain extra care by eliminating the host nestlings. It is not related to the foster parents, its demand for unlimited parental care does not put a genetic sibling at risk, and it loses no kin by overworking the foster parents; but if a foster parent dies, the young cuckoo starves (Gochfeld 1979, McLean and Griffin 1991, Briskie et al. 1994, Dearborn 1998, 1999, Payne et al. 1998, 2001, Payne and Payne 1998b, Kilner and Davies 1999, Davies 2000).

Effects of a brood parasite on its host Brood parasites reduce the nesting success of their hosts (Rothstein 1975b, Payne 1977b, 1997a). In the relatively benign brood-parasitic finches and cowbirds, the removal of a host egg by the female parasite accounts for most of the reduction in nesting success of the host (Morel 1973, Payne 1997a, Payne and Payne 1998b). In addition, brood-parasitic nestlings compete with host nestlings for parental care.The tactics of young parasites to control parental care for their own benefit are a counterpart of the sibling competition of certain nonparasitic nesting

birds.Within a brood of nesting birds, asynchronous hatching and not-so-subtle siblicide may be advantageous to the rival that wins the struggle with its siblings (Nathan et al. 2001).The young brood parasites carry the theme of sibling rivalry to its extreme. Most parasitic cuckoos have a more rapid incubation than their hosts and are larger at hatching. The head-start in development gives the young parasite an advantage in winning parental care (Payne 1977b). Second, some cuckoos are larger, grow more rapidly and take more food than their hosts, as in the Great Spotted Cuckoos. Third, brood-parasitic cuckoos are the only nestlings that regularly and actively push and remove the host eggs and nestlings out of the nest. Brood-parasitic New World cuckoos are equally deadly to their nestmates, killing the host young with their hooked bills. Although the female cuckoo causes some loss of host success when she removes an egg from the host clutch, the later eviction and biting behavior by the nestling cuckoo causes the total loss of the host brood. The frequency of brood parasitism affects the breeding success of the host population and the evolution of host defense against brood parasitism (Davies et al. 1996). The proportion of nests parasitized by Common Cuckoos is less than it was in earlier decades in Britain (Brooke and Davies 1987), yet in some areas of Europe the intensity of cuckoo parasitism remains high. There is often a great deal of regional and local variation in parasitism. Songbirds parasitized in one area are not parasitized in another area, even though cuckoos live in both areas (Makatsch 1937, 1955). In southern England where the regional rate of parasitism of Reed Warblers is about 5%, the local rate of cuckoo parasitism of warbler nests ranges from 0 to 16% (Lindholm 1999). In Central Europe where the rate of cuckoo parasitism per nest averages 8.3% for Reed Warblers and 6.3% for Marsh Warblers, from 3% to 28% of the warbler nests in local populations are parasitized. In Central Europe the most frequently parasitized host is the Great Reed Warbler, and in the marshes of Hungary nearly 50% of warbler nests were parasitized in the early 1940s (Molnár 1944).The rate of parasitism has remained high for decades, and surveys in the late 1990s showed a higher rate of parasitism (57%) with half

Brood parasitism 151 the parasitized nests having 2 or more cuckoo eggs (Moskát and Honza 2002). In Africa and Australia a high proportion of host nests are parasitized at least locally by several cuckoo species, though the proportion is much less than 50% (Payne and Payne 1967, Brooker and Brooker 1989a, Payne 1997a). The effect of parasitism on a host population can be calculated from field data. The proportion of nests that are parasitized, multiplied by the difference in success in parasitized and unparasitized nests (the number of host young that fledge, or the proportion of nests that fledge at least one host young), allows for not all unparasitized nests succeeding and not all losses of parasitized nests being due to parasitism (Payne 1977b, 1997a). Egg predation by cuckoos also reduces breeding success of the hosts, and in some populations more nests have eggs taken by a cuckoo than nests parasitized by the cuckoo (Schulze-Hagen 1992). Costs of Common Cuckoo parasitism to Reed Warblers were calculated from the breeding success of host pairs with their own young (Øien et al. 1998). In a population in the Czech Republic, 16.1% of 1108 warbler nests were parasitized; in this population the cuckoo eggs are not good mimics of Reed Warbler eggs. In parasitized nests the nesting pairs that removed the cuckoo egg had a success of 0.29 for each warbler egg, and pairs that accepted the egg had a success of only 0.03, yet the warblers removed the cuckoo egg from only 8% of the nests. Some birds deserted and nested again; 30% of parasitized nests and 1.8% of unparasitized nests were deserted.These observations suggest that warblers which accept a cuckoo egg in the nest do not behave in their own best interests. Success in the number of warbler young fledged from unparasitized nests was 1.95, and the number fledged from parasitized nests was only 0.11. When the effect of cuckoo parasitism on production of young in the host population is calculated as % parasitized nests x% difference in success of parasitized and unparasitized nests (the mean number of eggs in unparasitized nests ⫽ 4.22), the success of unparasitized nests was 1.95/4.22 ⫽ 0.462, and the success of parasitized nests was 0.11/4.22 ⫽ 0.026), the difference is 0.436, or 44%.That is, cuckoo parasitism was responsible for nearly half the loss in breeding success in the population. In nests where cuckoos

may have removed a host egg, the number of host eggs in unparasitized nests (3.88) and number in the parasitized nests (2.12) allows an assessment of costs that are accounted to the female cuckoo and to the nestling cuckoo. In this case, 1.76 eggs were lost to the cuckoo female, and the difference (1.95 young fledged in unparasitized versus 0.11 young in parasitized nests) was 1.84 young to the cuckoo nestling. That is, nearly half the cost of being parasitized was the result of egg removal by the cuckoo female, and more than half the cost was due to eviction of host eggs by the cuckoo nestling. When the cuckoo female removes a host egg or two from the nest, the host cannot recover the loss by removing the cuckoo egg; nevertheless a host could avoid the cost of nestling eviction by a cuckoo chick if the host removes the cuckoo egg. If the host were uncertain which egg is the cuckoo’s, the host could gain only 0.03%, the proportion of nests that would otherwise fail. For comparison, the cost of removing a cuckoo egg is more than 2%, the proportion of times it damages or removes its own egg when it attempts to remove a cuckoo egg (Davies et al. 1996).A host should remove a cuckoo egg when the cuckoo egg differs from its own egg, but not when it is uncertain which egg is the alien egg, as when the cuckoo egg is a good mimic. In fact, female warblers that lay variable eggs themselves are less likely to remove a cuckoo egg than females whose own eggs look alike. Perhaps the variability in their own eggs leads to uncertainty and inaction against the parasite egg (Stokke et al. 1999). The most complete records of the effects of cuckoo parasitism in the southern hemisphere are for hosts of the glossy cuckoos Chrysococcyx spp. (1) In New Zealand, Grey Gerygones are parasitized by Shining Bronze-cuckoos. Nests not parasitized had a 33% success in fledging at least one young, and nests with a cuckoo egg had only 1.9% success in fledging a young of the host. In this population 55% of 40 nests were parasitized (Gill 1983a).The impact of cuckoos on decreasing the production of host young was (0.55 ⫻ (32.9 ⫺ 1.9)/32.9) ⫽ 17% (Payne 1997b). (2) In Western Australia, Splendid Fairywrens are parasitized by Horsfield’s Bronze-cuckoos (Rowley et al. 1991).The incidence of brood parasitism was 20%. In some parasitized nests the cuckoo

152 The Cuckoos egg was laid before the host had laid and the fairywrens then deserted, and in others the cuckoo egg failed to hatch or survive, but parasitized nests failed completely when the cuckoo evicted the host eggs. Nests that were not parasitized had a mean success of 1.70 fledged fairy-wrens, and nests that were parasitized had 0.11 fledged fairy-wrens (Brooker and Brooker 1996). Because mean clutch size in unparasitized nests is 2.91, when the female cuckoo removes an egg and mean clutch size in parasitized nests is 1.89 host eggs, the cost of parasitism is due mainly to evicting behavior by the nestling cuckoo. The overall impact of cuckoos was to decrease the success of fairy-wrens by 19%. (3) In South Africa, Southern Masked Weavers are parasitized by Diederik Cuckoo. Nests not parasitized fledged on average 1.48 young weavers, and nests with a cuckoo egg fledged only 0.16 young weavers. In a few parasitized nests the cuckoo egg failed to hatch and the young weavers survived; in both kinds of nests some broods failed owing to storms and predators. In this population 43% of 76 nests were parasitized (Hunter 1961). For the host of Diederik Cuckoo, the impact of the cuckoo was to decrease the production of host young by 38%. A female cuckoo can make an impact on the breeding success of a local host population. In Britain, Chance (1922) found that a female Common Cuckoo that he knew on sight laid as many as 25 eggs in a season in nests of its host Meadow Pipit. Similar large numbers of eggs were laid by a female cuckoo in host nests in northeastern France (Blaise 1965). In southern Africa, female cuckoos lay 12–24 eggs in a season. Seasonal egg laying was determined by the proportion of females with an egg in the oviduct, the number of recently ovulated follicles in the ovaries, and the length of the local breeding season.A female Diederik Cuckoo lays 16–21 eggs in a season and a female Jacobin Cuckoo lays 19–25 eggs. Other species that have been examined (Great Spotted Cuckoo, Levaillant’s Cuckoo, Black Cuckoo, Redchested Cuckoo, Klaas’s Cuckoo, African Emerald Cuckoo) have similar rates of laying. Small cuckoos lay on successive days, while large cuckoos and even the small glossy cuckoos sometimes lay on one day, carry an egg in the oviduct for a day then lay on the following day, and do this in a series (or “clutch”) of 3–5 eggs in a set with a few days between the series.

Between sets the ovary has only small ovarian follicles, then it yolks up another series, staggered in size with the largest developing follicle about twice the size of the next one and so on. A female cuckoo can lay no more than one egg in a day. Like other birds the cuckoos have only one ovary and one oviduct, and their reproductive anatomy enforces the “egg-aday” limit in birds, even in chickens that are bred to lay many eggs. Ovaries of brood-parasitic cuckoos were compared with those of breeding Yellow-billed Cuckoos taken at nests where they incubated or had small nestlings, to calibrate how the ovarian follicles of cuckoos change from the time the egg is ovulated and laid, to the time the bird was on the nest (Payne 1973a, 1974).

Behavioral response by the nesting host to the cuckoo Nesting hosts respond to an adult cuckoo near their nest by mobbing, flying at it, striking it and calling loudly (Smith and Hosking 1955, Payne et al. 1985, Macdonald 1990, Duckworth 1991, 1997, Yoshino 1999, Hanmer 2001). Mobbing may drive a cuckoo from the nest, and a host may avoid parasitism of its nest if the cuckoo leaves before she lays. But if a host mobs when a cuckoo is not near the nest, the conspicuous behavior may help the cuckoo to locate the nest, so a host often limits its mobbing behavior to the nest site after the cuckoo has already located the nest. Active defense near the nest can drive away the cuckoo, and strong defense can kill the cuckoo. A Smooth-billed Ani (an occasional within-species brood parasite) was killed by other anis as she attempted to lay in an ani nest (Loflin 1982). A Common Cuckoo with an egg in the oviduct ready to lay was killed by a Great Reed Warbler near a warbler nest; the female was found freshly dead, floating in the water by the nest just after the warblers were excited around the nest ( Janisch 1954).A Common Cuckoo near a nest of Bull-headed Shrike Lanius bucephalus was photographed aftger being killed and plucked by the shrike (Yoshino 1999). And a Diederik Cuckoo was killed when a male Southern Masked Weaver defended its nest site, harried the cuckoo to the ground, pecked it on the head and continued to attack it on the ground;“Mr Hoffman then held the

Brood parasitism 153 cuckoo up whereupon the weaver attacked the dead cuckoo in his hand” (Lorber 1985). Most small nesting songbirds in Europe reject the eggs of other birds. Rejection of a cuckoo egg is the main line of defense against brood parasitism. In field experiments, when an egg unlike their own (a “cuckoo” egg) was placed into the nest, eight species always or almost always (⬎ 80%) rejected the unlike egg, 16 species often (20%–80%) rejected the egg, and ten species seldom or never (0%– ⬍ 20%) rejected it. In the ten accepters, five were hole-nesting songbirds such as tits and wrens that are not parasitized in natural conditions (Øien et al. 1995).Variation in acceptance and rejection of an egg is also affected by the time when the egg is laid in the nest. The nesting females more often reject when a cuckoo egg appears while the hosts are laying, than when an egg appears after they have begun to incubate (Moksnes et al. 1994, Palomino et al. 1998). The egg colors and patterns of cuckoos have evolved together with the responses of the intolerant host species. Cuckoos of the mimetic egg morphs later colonized other host species that are not as discriminating and accept a cuckoo egg that does not closely match their own. In Europe, several songbirds are parasitized by Common Cuckoos with blue eggs, but of these hosts, Redstart Phoenicurus phoenicurus (blue eggs) and Whinchat Saxicola rubetra (pale blue eggs, sometimes speckled rusty) are the most discriminating, and the nondiscriminating hosts such as Wheatear Oenanthe oenanthe (white eggs spotted brown) and Pied Flycatcher Ficedula hypoleuca (pale blue eggs, rarely with fine speckling) may have been parasitized later by the cuckoo lineages (Moksnes et al. 1995, Gibbs et al. 2000). The risk of removal or damage to their own egg may explain why some hosts accept a cuckoo egg in their nest. When a cuckoo egg is laid late or when it is sterile, there is no benefit in removing it, and the same hosts that remove a cuckoo egg when it is laid along with their own clutch will accept a cuckoo egg when it is laid well after incubation has begun. But when the hosts removes an egg, they risk removing or (perhaps more importantly) damaging their own egg (Marchetti 1992, Soler et al. 1996, Røskaft et al. 2002).

In contrast to the small hosts of Common Cuckoos in Europe which reject a cuckoo egg, small songbirds in North America generally accept the egg of the brood-parasitic Brown-headed Cowbird (Rothstein 1975a, 1990).The behavior of the nestling parasites in a parasitized nest accounts for this difference in the behavior of the nesting hosts. A young cuckoo kills the host eggs or nestlings; a young cowbird often grows up together with the host nestlings. In cuckoo hosts the main cost of being parasitized is paid after the cuckoo hatches, whereas in cowbird hosts the cost is paid when the female cowbird removes a host egg (Moksnes et al. 1991, 1994, Payne 1998, Øien et al. 1998, Payne and Payne 1998b).The greater impact of a nestling cuckoo on the nesting success of the host explains why cuckoo hosts go to great lengths to remove a cuckoo egg, even puncturing the eggshell and risking their own eggs. A young cuckoo looks different from the hosts’ own young, and it is amazing that the foster-parents rear these strange young in their nest. Observations of songbird hosts and of other brood-parasitic birds give only partial answers to the question of why the foster parents rear young that are not their own. Brood-parasitic finches Vidua have nestlings that mimic the nestlings of their foster parents’ own chicks with which they grow up together in the nest.The visual resemblance of nestling Vidua to chicks of the foster species may allow the parasite to be accepted and provided with parental care (Payne 1997a, 1998). Except for Thick-billed Cuckoos which have mouth spots and whose host bush-shrikes Prionops have mouth spots (the details are undescribed), no such visual mimicry is known in nestling cuckoos.Within a species, parents do not discriminate in caring for young that are genetically their own and young that were fathered by another bird or were laid in their nest by a conspecific female (Westneat and Sherman 1993, Westneat and Sargent 1996, Whittingham and Dunn 1998, Payne et al. 2001). Magpies are more likely to accept an experimentally added chick of the Great Spotted Cuckoo if their nest already has a nestling cuckoo, than if it is not parasitized (Soler et al. 1995a). However, the hosts of some brood parasites rear the young of other species even in a mixed-species brood where the nestlings differ in appearance and in begging calls (Payne et al. 2001).

11 The evolution of brood parasitism in cuckoos

Equilibrium and coevolution Although a cuckoo nestling destroys the host eggs and nestlings, the host parents often do not discriminate against a cuckoo egg and they incubate it with their own eggs.The hosts may lack the genetic variation that would allow evolutionary selection to favor action against cuckoos (Rothstein 1990). Or, the hosts and their parasitic cuckoos may be in a state of equilibrium, where the gain of rejection is balanced by the cost of rejection of the cuckoo egg (Davies and Brooke 1988, Davies et al. 1996). The costs and benefits of brood parasitism and response by the hosts to the parasite can be described as a mathematical model, which assumes genetic variation in both the parasite and host and tracks the coevolutionary outcome of these costs and benefits. These two hypotheses, the lack of genetic variation in hosts and the equilibrium in costs and benefits in response to cuckoo parasitism, differ in their ability to predict the variation in behavior among host birds (Payne 1997a). Genetic differences that cause behavioral differences or are heritable along family lineages have not been demonstrated in the field or in the genetics’ laboratory. Models that explore the effects of population size and the intensity of brood parasitism assume genetic differences between hosts that defend and hosts that accept the cuckoo egg, with “rejecter genes” in the host and “egg color genes” in the cuckoo, then track the success of the assumed gene under different ecological conditions (Kelly 1987, Takasu et al. 1993, Takasu 1998). Although there may have been genetic differences in the past, the status of this

model is one of indirect reasoning. In contrast, the ecological factors that determine the alternative behavior have been identified, and the costs and benefits of alternative behaviors to birds with a cuckoo egg in their nest have been measured in the field. In consequence we can describe the benefits to individual hosts that use these behaviors, in terms of the number of their young that survive the period of parental care.When the frequency of parasitism and the decrease in fitness of parasitized nests are high, we expect selection to favor hosts that remove the cuckoo egg. This ecological and genetic interaction is combined in the coevolution model. A balance or dynamic equilibrium in the ecological and genetic interaction between host and brood parasite accounts for the adaptations of host and cuckoo.The evolution of response of a host species to cuckoo parasitism depends on the hosts’ frequency of being parasitized, the costs of being parasitized and the costs of defense (Lotem et al. 1995, Davies et al. 1996). Individual hosts vary in their response to a cuckoo egg, and a host that removes an egg from its nest at one time may accept it another time. Whether a host rejects or accepts the cuckoo egg depends on variable conditions such as the host’s age and experience, the stage of nest, and the habitat, as will be seen. (1) The cost of removal may be greater than the benefit gained by removing the egg. Costs occur if the nesting bird fails to discriminate between its own egg and a cuckoo egg, and removes its own egg in error; or if it damages its own egg when it attempts to remove a cuckoo egg (Røskaft et al. 1993, Lotem et al. 1995, Davies et al. 1996). (2) The numbers of cuckoos in a nesting area

The evolution of brood parasitism in cuckoos 155 may affect whether a host removes the egg (Davies et al. 1996). Reed Warblers have a higher rate of rejection of model cuckoo eggs in parasitized than in unparasitized populations (Lindholm and Thomas 2000). Over a 12-year period from the 1980s to the late 1990s in one population, the rate of cuckoo parasitism declined from 20% to about 5%, along with a decline in the number of cuckoos. In experiments with model eggs in warbler nests, there was a decline in host rejection of nonmimetic eggs, from 75% in 1985–1986 to 25% in 1997, and a decline within a season, with the host less likely to reject when cuckoos disappear (Brooke et al. 1998, Lindholm 2000). The decline in the rate of rejection was too rapid to be owing to genetic change, and it may have been in response to the changed environmental conditions such as the number of cuckoos in the breeding area. (3) In field experiments, nesting birds were likely to reject the egg when an adult Common Cuckoo dummy was near the nest, but to accept it when there was no cuckoo dummy (Davies and Brooke 1988, Moksnes and Røskaft 1989, Moksnes et al. 1990, 1993a, Lerkelund et al. 1993, Alvarez 1996, Davies et al. 1996). In tests with another host and parasite, Magpies were no more likely to reject a model egg when a hand-reared Great Spotted Cuckoo perched near the nest than when the nest had no cuckoo nearby (Soler et al. 2000).This host is more likely to accept a cuckoo egg than are hosts of the Common Cuckoo, because nestling Great Spotted Cuckoos do not evict the host eggs. (4) The stage of breeding affects rejection, as birds are more likely to reject a cuckoo egg during laying than during incubation, when a late cuckoo egg would not hatch before the host eggs hatch (Moksnes et al. 1990, 1993a). (5) The number of eggs in the nest may affect whether a host deserts the nest as happens when the cuckoo takes more than one egg, or when a predator reduces the clutch size (Øien et al. 1998). (6) Experience with her own eggs may affect whether a nesting host rejects an egg in a later nest. In experiments that introduced dummy eggs into the nests of Great Reed Warbler, the yearling females were more likely to accept an egg than were the older females (Lotem et al. 1992). (7) Experience in rearing a cuckoo may determine the

host’s behavior in a later nesting. Magpies that have reared a young Great Spotted Cuckoo are more likely to accept another cuckoo if it is experimentally placed into the nest, compared to adults that have not reared a cuckoo. When Magpie and cuckoo nestlings are reared together, the older they are the more the Magpie parents discriminate between them. When a fledged cuckoo is placed near their nest, they feed their own young in preference to the young cuckoo, and they tolerate alien Magpie nestlings more often than cuckoo nestlings (Soler et al. 1995a). Although the host young often survive and leave the nest, the young cuckoo takes its share of parental care and food and affects the growth and development of the young Magpies (Soler et al. 1996). Cuckoo fledglings are loud and persist in begging, they attract other birds and these newcomers may feed the cuckoos: as many as four species have fed a fledgling cuckoo (Kikkawa and Dwyer 1962, McBride 1984,Yoshino 1999. In summary, the response of hosts to a cuckoo egg varies with the risk of being parasitized, the cost of accepting the cuckoo egg, and the gain to be realized by rejecting the cuckoo egg. A conditional response explains why some hosts accept the cuckoo egg at one time, but reject it at another time (Lotem et al. 1992, Lotem and Nakamura 1998, Soler et al. 2000). A conditional response can be tested with predictions in experiments and in observations of long-term changes in the behavior of a bird (Davies et al. 1996). Coevolution is the genetic evolution of adaptations between interacting species. This idea describes this interaction over evolutionary time between brood parasites and their hosts, especially where the interaction is reciprocal with each responding to the adaptations of the other. Brood parasites that lay eggs with the color and pattern of the host are less likely to have their eggs rejected, and the evolution of similar eggs is a response to selection by their hosts (Baker 1923, 1942; Payne 1977a; Brooke and Davies 1988; Davies and Brooke 1989a,b; Rothstein 1990). The hosts use both color and spot differences between a cuckoo egg (or a model egg) and their own egg; and the more similar the alien egg is to their own, the more likely it is to be accepted by the host, as in

156 The Cuckoos Village Weavers in West Africa (Lahti and Lahti 2002). Parasitized species are more variable in egg color and pattern than are unparasitized species, and the different levels of variation suggest that egg color and pattern evolve in response to cuckoos (Øien et al. 1995, Soler and Møller 1996). The critical test is to determine whether the egg color and pattern change through time and in what direction they change. The behavior of nesting birds may have evolved towards the mismatched eggs in their nest in response to cuckoos, but their behavior also has other explanations. Some birds may reject in response to an individual bird’s experience and history of intraspecific parasitism (Brown and Brown 1989, Westneat and Sherman 1993).The variability in egg color between females in some colonial weavers may have evolved to allow them to distinguish their own eggs from those of other female weavers, rather than from eggs of cuckoos, which parasitize the weavers infrequently (Freeman 1988, Jackson 1998). More often, many birds remove an egg in their nest if it is damaged; and the removal of an egg that looks different from others in the clutch may increase the success of the remaining eggs (Rothstein 1990). In the case of antagonistic coevolution, when the mimic gains an advantage in resembling another species, its model, the model may be at a disadvantage.The model species may be selected to escape this mimicry, in a process of chase-away selection (Fisher 1930, Gavrilets and Hastings 1998). Fisher (1930) noted that a model might change its signals to escape its mimic. On the other hand, if a genetic variant in the model species produces a different egg color and pattern from the egg-mimetic cuckoo, then the model might be challenged to recognize its own eggs, because in changing its characters it might not identify its own eggs; and the disadvantage might be greater than the advantage gained in being different from the mimic (Nur 1980).Turner ( 1984: 154) argued that “The advantage of being a mimic…is considerably greater than the disadvantage of being a model”, and he questioned whether a model would diverge from its mimic (Turner 1995). In at least some cuckoo hosts, a female recognizes her own eggs in the nest, by learning their appearance when she lays

her first eggs, accepting them and rejecting eggs of different appearance that may appear late in the nestr (Lotem et al. 1995). This development of behavior provides a mechanism to escape the challenge of egg mimicry and to coevolve along with the egg-mimetic cuckoo. Both a coevolution model and an equilibrium model predict that nesting birds with a long history of being parasitized are more likely to reject a cuckoo egg than birds in an unparasitized population. Indeed, Meadow Pipits are more likely to accept nonmimetic model eggs in Iceland where cuckoos do not now occur, than in Britain (Davies and Brooke 1989a). Pipits also accept such eggs in Norway, where they were derived from a population that was parasitized but where the probability of being parasitized is low (Moksnes et al. 1993). Both a coevolution model and an equilibrium model also predict that the variation in color and pattern of eggs among clutches of the host species is greater in parasitized birds than in unparasitized birds. The eggs of European songbird hosts are in fact more variable than are the eggs of unparasitized songbirds (Moksnes and Røskaft 1995). At the same time, the variation among eggs within a clutch is lower in the parasitized host species that remove a cuckoo egg (Øien et al. 1995, Soler and Møller 1996). It appears there has been selection within a host species to evolve eggs they can recognize, and to reject the cuckoo egg. The coevolution model and the equilibrium model lead to the same predictions, and they are not mutually exclusive. One concerns an evolutionary time scale.The second can be examined directly and tested experimentally on a behavioral time scale. In practice, the hypothesis of lack of genetic variation is a null hypothesis used by default when other hypotheses fail to show a difference in the behavior of individuals within a population.The null hypothesis has been applied in comparisons between species, rather than within a species at the level where natural selection is effective (Rothstein 1975a, 1990).

Have cuckoos and hosts changed their behavior in historical time? Brood parasitism in cuckoos sets a wide stage to consider interactions between species, and different

The evolution of brood parasitism in cuckoos 157 acts are in progress on this stage. A change in the behavior of hosts and parasites over the years might indicate natural selection of one species (the parasite or the host), and an interaction between parasite and hosts; cuckoos have changed their host species in some regions (Baker 1942, Nakamura 1990). In Burma near Mogok, Maymyo and Taunggyi, Common Cuckoos changed their hostchoice behavior during the early twentieth century (Osmaston 1916, Livesey 1933, 1933, 1936, 1938a, Baker 1940, 1942). Hosts that were once common became scarce as the habitat changed from forest to open country, and the most common cuckoo egg changed at the same time. In the early twentieth century (1900–1915) the common host was the Grey Bushchat Saxicola ferrea, which laid blue eggs. In the 1930s the common host was the Stonechat Saxicola caprata, which laid pink eggs with reddish spots. In the early years cuckoos parasitized the common Grey Bushchat, and cuckoo eggs matched the blue bushchat eggs. In later years, with increased cultivation of rice, gardens and village vegetation that replaced the scrub forest, stonechats replaced bushchats, and the most common cuckoo egg morphs in the region were pink with reddish spots, like the stonechat eggs (Baker 1940, 1942). Egg colors are genetically determined, and a change in the frequency of color morphs of these cuckoo eggs indicates a change in the genetic makeup within the cuckoo population through natural selection. In this case the change in cuckoo eggs occurred when habitat change favored the Stonechat and the egg race of the cuckoo that mimicked the new common Stonechat host. In Britain, nest records of songbirds compiled over 40 years for the six main hosts of the Common Cuckoo include a remarkable sample of 73,750 nests, with 1,016 nests parasitized by cuckoos (Brooke and Davies 1987). During this period there were decreases in the proportion of parasitized nests in the Dunnock, European Robin and Pied Wagtail. For the Reed Warbler there was a small increase in the proportion of nests parasitized. No change appeared in the accuracy of cuckoo egg mimicry of Reed Warbler eggs, and no cuckoo eggs appeared that mimicked Dunnock eggs. In this population, the change in the cuckoo’s use of host species was

not accompanied by a change in cuckoo egg mimicry or in host behavior (Brooke and Davies 1987, 1998; Davies and Brooke 1988, 1989a,b). In three case histories a change in behavior has been reported in the host, the cuckoo, or both host and cuckoo.These cases have been considered natural experiments in evolution (Davies 2000). Individual experiences of the host or the cuckoo can also explain the observed changes. (1) Village Weavers in Africa are parasitized by Diederik Cuckoos, and in field experiments in West Africa these weavers remove from their nest any eggs unlike their own (Lahti and Lahti 2002). In the West Indies, where weavers were introduced into Hispaniola 200 years ago, there are no broodparasitic cuckoos. In a field experiment in Hispaniola the nesting weavers did not discriminate against eggs unlike their own (Cruz and Wiley 1989); in a subsequent test the weavers did remove eggs unlike their own (Robert and Sorci 1999).The lack of rejection in the early test suggested that when there were no cuckoos to parasitize the weavers, rejection was not selected, perhaps because rejection was costly as the birds removed their own eggs. However that study did not carry out a control test in Africa to determine whether the lack of response was due to the experimental method. Hispaniolan weavers all have blue eggs; in contrast, within a local population in Africa, egg color and pattern is highly variable (Collias and Collias 1959, Maclean 1993, Lahti and Lahti 2002).The low range of variation in the model eggs, which were mainly the blue eggs of local weavers in Hispaniola, may explain the results of the early test, while in the subsequent test the model eggs were more variable and the results were similar to the tests in West Africa. (2) Magpies in Mediterranean Europe are parasitized by the Great Spotted Cuckoo. Magpies sometimes reject a model cuckoo egg from their nest, but in northern Europe where the cuckoo does not occur, the Magpies do not reject the model egg. The difference suggests that Magpies have changed their behavior in response to the risk of cuckoo parasitism, either where they are parasitized or even where they are not parasitized. Great Spotted Cuckoos have been studied most intensively by a research group at the University of

158 The Cuckoos Granada. Soler and Møller (1990) found a difference in Magpie behavior in three sites, two in Spain and one in Sweden. Magpies were more discriminating in Spain at Santa Fe, less so 60 km from there at Guadix, and not at all in Sweden where the cuckoos do not occur. The research group suggested that the longer coexistence of host and cuckoo explained the difference between Santa Fe and Guadix (cuckoos were not reported at Guadix before 1962).The results were repeated (Soler et al. 1994b) and the authors interpreted the difference between sites as “micro-evolution”.This interpretation has been questioned. The sites were only a day’s walk apart on a flat road, and their report of differences in the time when cuckoos and Magpies lived together were based on conversations with hunters.There was no direct observation of change in Magpie behavior or cuckoo parasitism over time within a population (Zuñiga and Redondo 1992a). The difference in host behavior between sites could also be explained by the numbers of cuckoos present at each site. In accord with the authors’ interpretation, however, when a tame cuckoo was perched near the nest, there was no increase in rejection rates (Soler et al. 2000). (3) The third case involves a reported change in both host and cuckoo. In Nagano Prefecture in Japan, the Common Cuckoo first appeared in 1960 as a parasite of the Azure-winged Magpie Cyanopica cyana. Cuckoo parasitism increased in the 1960s and 1970s when the altitudinal ranges of cuckoo and magpie expanded and the birds came into contact. Cuckoo parasitism increased from 0 to 30%, then it decreased, and at about this time the magpies rejected cuckoo eggs (Yamagishi and Fujioka 1986, Nakamura 1990). Observations suggested that cuckoos had parasitized magpie nests for a different length of time in three localities: 20 years at (a) Azumino, 15 years at (b) Nagano, and 10 years at (c) Nobeyama. When host magpies were tested with model cuckoo eggs, in (a) and (b) they rejected eggs at about the same rate. Habitat, breeding density and behavior differed among the cuckoo populations, and the comparison does not necessarily reflect a change in parasitism through time, although the field workers interpreted the results in this way (Nakamura et al. 1998).

Cuckoo eggs were compared to see whether cuckoos changed genetically to mimic the host species’ eggs. Cuckoo eggs were sampled in the localities where magpies were tested for their response to model cuckoo eggs. Eight cuckoo egg morphs were defined by lines, dots and blotches, with five egg morphs in magpie nests (Nakamura et al. 1998). There were no statistically significant differences between populations when all egg morphs were considered together, and at Nagano the three host species all had the same most common cuckoo egg morph in their nests.The authors concluded that the different proportions of cuckoo egg morph “A” between (a) Nagano and (c) Nobeyama resulted from the longer history of cuckoo parasitism at Nagano. Magpie eggs have dots and blotches but no lines, unlike the cuckoo eggs in their nests. Because there was no precise mimicry in cuckoo eggs, no shift in the makeup of eggs in the region through time, no overall significant difference in the frequency of cuckoo egg morphs between localities, and no higher proportion of mimetic cuckoo egg morphs in magpie nests in the locality with the longest known history of cuckoo parasitism, there is no compelling evidence to support the conclusion that cuckoos changed by natural selection within two decades to match the eggs of a new host species. Magpies banded at Nagano dispersed within their breeding area and were not seen outside the area (Hosono 1969). Local studies underestimate dispersal (Payne 1990), and it is likely that magpies born outside the study area moved into the population and cuckoos moved as well. Azure-winged Magpies are parasitized elsewhere, in China by the Indian Cuckoo (Shaw 1938, Hoffman 1950) and in Spain by the Great Spotted Cuckoo (Arias de Reyna 1998). The magpie response to cuckoo parasitism may have evolved in other times and places. Azurewinged Magpies are known from fossils dated to 44,000 years ago in the Iberian peninsula (Cooper 2000), and magpies have had a long time to respond to cuckoos.Any changes in behavior of cuckoos and magpies in Japan may not have involved natural selection, insofar as a switch in host acceptance of a cuckoo egg might involve their own experience rather than genetic change within a population.

The evolution of brood parasitism in cuckoos 159 In summary, none of the three cases demonstrate a change in the host, in response to the onset and progress of brood parasitism through time, and none show a change in the eggs or the parasitic behavior of the cuckoos more clearly than the observations in Burma in the early 1900s.

Adaptation and the evolution of brood parasitism The parasitic cuckoos are “obligate” brood parasites. Within a species, all females lay their eggs in the nest of another species of bird and never nest themselves; they are committed to this behavior and are constrained by the lack of any nestbuilding behavior or parental care, to depend on other species to incubate their eggs and to brood and rear their young. The origin and evolution of this brood-parasitic life style has two sets of explanations. One involves the adaptive ecology of parental care and dispersion of the eggs to avoid predation, and a second compares life styles among birds and the phylogenetic relationships of brood-parasitic birds to nesting birds (Hamilton and Orians 1965, Payne 1997a).

Adaptations and behavioral ecology Nesting birds are attentive parents, incubating their eggs, then bringing food to their young until they are independent, and they show the same behavior to the parasitic cuckoos (Lack 1954, Clutton-Brock 1991). In contrast the brood parasites depend on other species to hatch and care for their young. Studies on parental care of breeding birds define the limits of rearing their own young and the consequences of laying in the nests of other birds. First, caring parents may be limited in the amount of food they can find and provide to their brood (Lack 1954, 1968). Second, birds that lay additional eggs in the nests of other birds that provide parental care can leave more offspring in a season than a bird that lays only in its own nest (Darwin 1859). Interspecific brood parasitism in the cuckoos may have originated from within-species parasitism that produces two advantages: overcoming the limits of

parental care, and avoiding predation and minimizing the risk of failure.

Risk of predation Predation at the nest is a major limit to success in breeding birds. Birds that lay all their eggs in one nest are at risk of losing the entire brood. Laying eggs in the nests of other birds can spread the risk and increase the chance that at least some young survive. This diversification of parental effort is a case of “spreading the risk” or “bet hedging” (Seger and Brockman 1987, Brown and Brown 1996) and in particular it is a case of “risk aversion” by spatially dispersing the reproductive effort (Slatkin 1974)— much as some birds may be risk averse in nesting repeatedly, each time with a small clutch.When there is a high risk of nest predation (the chance of losing a nest to a predator is as high as 0.9 in some regions, Ricklefs 1969), dispersing the risk across several nests may increase reproductive success. The risk of nest predation may have led brood-parasitic birds to put their eggs up for adoption in other nests, and not “all their eggs in one basket” (Payne 1977b,c). Consider a female that either lays 4 eggs in one nest or lays them in several nests. Females that lay all the eggs in one nest have a success of 0 or 4, with the rate of nest predation determining the mean success. Females that lay each egg in a different nest have a reproductive success (RS) of 0, 1, 2, 3 or 4, and the proportion that have 0 success will be less than in females that lay all the eggs in one nest.The probability that all eggs will survive is the last term in expansion of a binomial (p ⫹ q)n ⫽ 1, where p is the probability of predation, q is the probability of not being taken by a predator, and n is the number of nests in which eggs are laid. Insofar as predators usually take all the eggs and young in a nest, the probability that at least one egg will survive is the same as the survival of at least one nest with the egg, 1 ⫺ pn. If predation risk p is high, then the chance that at least one egg of a female that lays her eggs in several nests will survive is higher than that of a female that lays all her eggs in one nest. In an analytic solution based on field data of risk, then p ⫽ 0.8 and q ⫽ 0.2, the chance that at least 1 egg in 4 survives, 1 ⫺ p4, is (1⫺0.41), or 0.59, higher

160 The Cuckoos than when all eggs are in the same nest q, or 0.20. The proportion of females with success of 4 (n, when all eggs survive) is less when females lay each egg in a separate nest, qn ⫽ (0.2)4, or 0.0016, than when females lay all their eggs in one nest, q1 ⫽ 0.2. Within a population the RS of females laying “all eggs in one nest” will vary, high in some females and low in others. In a large population the average success of females with all eggs in one nest is the same as that of females that spread the risk across nests. Although the likelihood of an individual female leaving any offspring is higher when she lays her eggs in the nests of other birds, her expected fitness is lower than in a female that has a successful nest with a full set of her own eggs. Laying eggs in the nest of another bird may spread the risk against nest predation, but eggs in the nest of another species are at risk of discrimination and rejection by the nesting birds. If the probability of rejection (i.e., infanticide) is high, then the better strategy may be to not subject all the offspring to adoption. Alternatives are to lay from 1 to 4 eggs in the bird’s own nest and the remaining eggs in other nests, in mixed laying strategies. If p is the probability of predation and r is the probability of rejection, then the expected number of surviving offspring in each strategy would be the expanded form of the equation n(1⫺p⫺(1⫺p)r)^n ⫹((n⫺1)(1⫺p⫺(1⫺p)r)^(n⫺1))(p⫹(1⫺p)r) ⫹ (( n ⫺ 2 ) ( 1 ⫺ p ⫺ ( 1 ⫺ p ) r ) ^ ( n ⫺ 2 )) ⫻(p⫹(1⫺p)r)^2 ⫹ (( n ⫺ 3 ) ( 1 ⫺ p ⫺ ( 1 ⫺ p ) r ) ^ ( n ⫺ 3 )) ⫻(p⫹(1⫺p)r)^3 . . . (I thank Charles Rosa for this formulation). This model refers to the probability of rejection, given no predation; that is, r is a conditional probability. In addition, the value of r is less when the bird lays all her eggs in her own nest than when she lays in the nests of other females. For example, with a risk of predation of 0.8 and a risk of rejection of 0.7, scattering all the eggs would be only marginally more successful than a mixed strategy of laying

some eggs in the nests of other birds and rearing the other eggs herself. If r ⬎ p, then the expected number of survivors in the second term of the equation could drop below the first term, “all eggs in one basket”. The dispersion strategy of one egg in each nest occurs at the value of r that makes the above expression ⬍ n(1 ⫺ p). In addition to risk of rejection, there is a cost in searching for nests, especially if the female also cares for a nest of her own. The trade-off between predation and rejection is somewhat like an investment problem where putting some resources out of one’s direct control reduces the chance of total loss, but also has a risk not present in an undiversified folio.The model of risk aversion is sensitive to both r and p. Consistent with the theory of risk management, when values were determined in cuckoos and their hosts, the observed values of egg rejection r were less than the values of nest predation p (Russell and Rowley 1993, Davies et al. 1996, Stokke et al. 1996, Rowley and Russell 1997). Regional variation in nest predation and the number of brood-parasitic species allow a partial test of the idea of brood parasitism as a strategy to minimize risk of loss. Predation risk on nests is high in the regions where brood parasitism is common. In the northern temperate region, fewer than half the bird nests are taken by a predator, whereas in tropical forest regions nearly 90% of nests are taken by a predator (Snow 1962, Ricklefs 1969, Willis 1973, Barnard and Markus 1990, Masuda and Ramanampamonjy 1996). Brood parasites are more diverse in tropical regions where the risk of nest predation is high. In northern temperate regions, only one species (Brown-headed Cowbird) is a brood parasite in most of North America; and whereas seven cuckoos (Hierococcyx hyperythrus, H. nisicolor, Cuculus micropterus, C. poliocephalus, C. canorus, C. optatus, C. saturatus) are brood parasites in the Old World, no more than four occur in a local site. In temperate Argentina there are three cowbirds and three cuckoos, and nearly the same number of brood parasites occur in the New World tropics with four cowbirds and three cuckoos. In tropical Africa there are 20 brood-parasitic finches, 14 brood-parasitic cuckoos, and 15 brood-parasitic

The evolution of brood parasitism in cuckoos 161 honeyguides (Sibley and Monroe 1990, Payne 1996). In tropical Asia the number of brood-parasitic cuckoos within a local region is as high as 13 along with a parasitic honeyguide (Malay Peninsula:Wells 1999). Although there are more species of birds in the tropics than in temperate regions, the number and diversity of brood parasites in the Old World tropics may be disproportionate to the total number of bird species in that region.

Cost of reproduction Even though the risk of nest predation may affect the evolution of brood parasitism, the extra parental care given to their young is the most conspicuous feature of reproductive behavior in altricial birds. Obligate brood parasitism is more common in altricial birds where it has evolved independently in several different families, but only once in the precocial birds where there is no extended parental care after the young has hatched (Sorenson and Payne 2001, 2002).The dependence of the altricial young through a long period is a major reason that cuckoos have evolved brood parasitism to exploit the parental care of other birds, because if the number of young produced is limited by parental care that the parents can provide, then altricial birds can lay more eggs and produce more young when they have at least some of their young reared by other birds (Payne 1965, 1977b, Lyon and Eadie 1991). The parental care provided by another species allows a female brood parasite to lay more eggs and leave more chicks than the bird could provide with its own parental care, much as Darwin (1859) suggested for a facultative brood parasite that lays some eggs in the nests of other birds, and rears other eggs the chicks of which it feeds in its own nest.This reasoning implies that reproduction is not constrained by a set number of eggs, as it is in the risk model. Rather, there is a trade-off between reproductive costs due to forming the eggs and costs due to parental care, so with no parental care, the brood-parasitic birds are expected to lay more eggs.When the number of eggs is greater than the difference between the expected survival of eggs in

a bird’s own nest and the survival of eggs in other nests, the brood parasite can beat the odds of predation as well as the costs of reproduction. In a test of this idea, brood-parasitic birds including the cuckoos lay more eggs than their nesting relatives (Payne 1974, 1977b,c). Further, when siblicidal behaviors that increase fitness by gaining parental care are selected in a facultative brood parasite, a mixed-strategy of facultative brood parasitism is expected to be less successful than a strategy of obligate brood parasitism. Once a lineage laid in the nests of other birds, aggressive behavior of these nestlings could favor the lineage that attacked or removed the nestmates (Roche and Glanz 1998). Survival of chicks in a nest is lower when more than one such supercompetitor cuckoo is in the nest (Robert and Sorci 2001). That is, obligate brood parasitism can be selected rapidly once the young are siblicidal, as they are in many obligate brood-parasitic cuckoos. The evolution of brood parasitism involves the number of eggs laid in pure and mixed laying strategies, the success of young in a nest with no other cuckoo and with other competitive cuckoos, and the costs of parental care in estimating optimal reproductive strategies.

Phylogeny and behavior models of brood parasitism Three models of breeding behavior Three life styles may have been involved in the initial evolution of brood parasitism. Each reproductive style could give an advantage to an opportunistic female, when her young gained extra-parental care as well as care by the breeding female (Payne 1977a, 1998). First, brood parasitism may have evolved from occasional laying in the nests of others of her own species. Second, brood parasitism may have involved frequent laying in the nest of cooperative breeding conspecifics such as in the communally nesting anis. Third, brood parasitism may have involved birds taking over the nest of other birds, first to rear their own young and then to leave their young to the care of

162 The Cuckoos the nesting bird. Each of these behaviors have been observed in bird populations, including some species that are related to the brood-parasitic cuckoos. To infer the likely historical origin of brood parasitism in a lineage of the cuckoos, these three summary models of behavior were compared among birds. Then the phylogenetic relationships of the brood parasites and their nesting relatives were compared, to determine which of these breeding behaviors occur in the lineages of the nesting birds that are most closely related to the brood-parasitic birds.

Intraspecific brood parasitism Obligate brood parasitism may have evolved through a mixed or conditional reproductive strategy (Payne 1977b,c) of birds to rear their own young and to lay an egg in the nests of neighbors. If these females laid more eggs, and these eggs were hatched and reared by neighbors while the errant females also reared their own brood, then they would achieve a greater reproductive success than females that laid only in their own nest and reared their own young, and the behavior would be selected. Black-billed Cuckoos and Yellow-billed Cuckoos usually rear their own young, but sometimes they lay eggs in other nests of their own species or another cuckoo’s nest (Fleischer et al. 1985). Darwin (1859) was aware of these nesting cuckoos’ behavior. He reasoned, “if the old bird profited by this occasional habit”, then laying eggs in a neighbor’s nest could lead to more offspring and to the evolution of brood parasitism. Facultative brood parasitism in nesting cuckoos suggests an adaptive precursor to obligate brood parasitism. Nestling Black-billed and Yellow-billed cuckoos have the same pattern of spots on the palate and tongue, displayed when they beg from the parents (Plate 20a,b). These patterns are likely to be traits of their common ancestor, rather than one species being a nestling mimic of the other. Facultative brood parasitism is common only in the precocial waterfowl. This behavior in the precocial birds argues against the importance of facultative brood parasitism as a necessary precursor for the evolution of obligate brood parasitism in altricial birds such as the cuckoos. In waterfowl, this

facultative parasitism involves the visual stimulus of an early nesting species such as Canvasback Aythya valisineria for the later nesting birds such as Redhead A. americana (Sorenson 1991, 1993, 1997), or nests available in which to lay when their own nests fail in Snow Goose Chen caerulescens that nest close to neighbors (Cooke 1987), or birds that lived together in their natal year (Andersson and Åhlund 2000). In some cases a female may specialize in laying in the nest of another. But in most field observations, a female’s decision to lay in the nest of another is related to her own condition or her recent nesting failure. A female whose nest is taken by a predator is likely to lay a few eggs in the nest of a neighbor, and a young female that arrives early when other ducks are nesting is likely to lay some eggs in the nest of another duck before she lays in her own nest.These females would leave more offspring on average than females that laid only in their own nest (Sorenson 1991, Åhlund and Andersson 2001), and it would be of interest to compare the seasonal success of the individual birds. The same idea applies to altricial birds. Facultative brood parasitism is known in some altricial species, especially birds nesting in dense colonies such as the Cliff Swallow Hirundo pyrrhonota (Brown 1984, Brown and Brown 1988, 1989, 1996, 1998) and Ploceus weaver finches (Dhindsa 1983, 1990, Jackson 1998).

Cooperative breeding Interspecific brood parasitism may originate with cooperative breeding.The cooperatively nesting anis and Guira Cuckoo live in social groups where several pairs share a nest. One female may be socially dominant over the others in an ani group, at least within a nesting cycle, and she tosses some eggs of other females from the nest before she lays. She and her mate incubate the clutch, while other pairs continue to incubate, brood and feed the young, some of them perhaps their own (Vehrencamp et al. 1986, Koford et al. 1990, Macedo 1992, 1994, Quinn and Startek-Foote 2000). Guira Cuckoos have laid in the nest of Smooth-billed Anis (Azara 1809). Parallels in the breeding of these crotophagine cuckoos with the breeding of female paper wasps

The evolution of brood parasitism in cuckoos 163 Polistes which fight over a nest, some winning and others abandoning, suggested to Hamilton (1964) the source of brood parasitism in cuckoos, as in Guira Cuckoos which remove the eggs of other females in their nest, and the altruistic behavior of guiras and anis in rearing any young in the nest. In this model, shared nesting provides a context for a female to lay in the nest of a conspecific. An unresolved problem is to shift the parental care to another species while retaining an advantage to the individual that is a parasite.

Nest takeover A bird may take over the nest of another species, then lay its eggs and care for its own young. In mixed-species colonies of weavers in Africa, Chestnut Sparrows Passer eminibey usurp other birds’ nests, though at times they build their own nests.The sparrows fight with weavers for access to fresh weaver nests, then lay their eggs and rear their own young (Betts 1966, Payne 1969c). In Kenya I watched the sparrows display persistently at fresh green nests that were built by other species, including Grey-capped Social-weaver Pseudonigrita arnaudi at Olorgesailie in 1967 and 1976, Goldenbacked Weaver Ploceus jacksoni and Black-headed Weaver at Kisumu in 1976, and Cardinal Quelea Quelea cardinalis at Kacheliba in 1976, and even at fresh, dry stick nests of White-billed Buffalo-weaver Bubalornis albirostris at Kacheliba in 1976.The sparrows displayed on the nests, attracted females, copulated on the nests, and entered the nests, the males first then the females. Both males and females brought a few nest materials into the nest and lined the nest. The nestbuilding birds chased away the sparrows, but the sparrows were persistent and they returned and displayed until the weavers deserted. Sparrows took over the nests of the other species; they were unsuccessful with the much larger buffalo-weavers. In other mixed-species finch colonies, some estrildid finches (Cut-throat Finch Amadina fasciata, African Silverbill Euodice cantans and Indian Silverbill E. malabarica) do the same, appropriating active nests, taking over old nests of weavers, or building their own nests (van Someren and van Someren 1945, Goodwin 1982, Sorenson

and Payne 2001). If a sparrow or finch were to leave its own egg when the nestbuilder is laying, rather than driving off the nestbuilder and appropriating the nest, then the weavers might rear the sparrow or finch young and these young would gain extra parental care from the weavers. The situation calls first for a mixed-species interaction (perhaps laying in other nests of their own species or sharing a nest between females, as in the silverbills), then a change in the behavior of the candidate parasitic bird to leave its eggs in the nest for the nestbuilding pair to incubate and rear, rather than driving away its potential host (Payne 1977b).

Phylogeny and the lineages of brood-parasitic birds In the larger view, it is likely that nesting evolved once in modern birds then was lost independently within a few lineages, three times in the cuckoos and four times in other brood-parasitic birds (once each in the waterfowl Anatidae, the honeyguides Indicatoridae, the parasitic finches Viduidae, and the parasitic cowbirds Icteridae, Payne 1977b, Sorenson and Payne 2001, 2002), and was lost in birds that breed in unusual ecological contexts such as penguins Spheniscidae that incubate their eggs on their feet (Williams 1995) and megapodes Megapodiidae that bury their eggs in the ground and leave their eggs to hatch with decomposing organic, solar or geothermal sources of heat, into precocial waifs without parental care ( Jones et al. 1995). The phylogenies uncovered in the mitochondrial genetic analyses of the cuckoos give an estimate of the number of times that cuckoos have evolved brood parasitism and the behaviors of the closest relatives of each of these parasitic lineages. Information on the phylogeny of brood parasites and their nesting relatives allows an inference of the breeding behavior of their most recent common ancestor. Using this line of reasoning, a variation on the theme of nest takeover was proposed for the origin of brood parasitism in cowbirds (Friedmann 1929). The Bay-winged Cowbird “Molothrus” badius takes over the nests of other birds, then lays its eggs and rears its own young in the

164 The Cuckoos nest (Hudson 1874, Sclater and Hudson 1888). The Bay-winged Cowbird in turn is parasitized by an obligate brood parasite, the Shining Cowbird Molothrus bonariensis. Current estimates of the evolution of icterids suggest that the nesting cowbird is not closely related to the brood-parasitic cowbirds: Bay-wing is now recognized in another genus, Oreopsar badius (Johnson and Lanyon 1999, Lanyon and Omland 1999). On the other hand, the birds most closely related to the brood-parasitic finches Viduidae are the waxbills Estrildidae. Brood parasitism evolved only once in the finches, including the cuckoo-finch Anomalospiza imberbis as well as the Vidua whydahs and indigobirds. Many species of estrildids appropriate the old nest of other nesting birds rather than building their own nest, but only a few are known to lay their eggs in the active nest of other species and none are cooperative breeders. The phylogeny of the Old World finches suggests that nest appropriation is the most likely behavioral origin of brood parasitism in the viduid finches (Sorenson and Payne 2001).The brood-parasitic honeyguides Indicatoridae are most closely related either to woodpeckers Picidae or to barbets Capitonidae (Short and Horne 2001). There have been no genetic sequence data published to resolve the relationship among these three groups. If the honeyguides are more closely related to the barbets, as suggested by Short and Horne (2001), then the nest appropriation behavior of the barbets in taking over old holes of other hole-nesting birds may have been involved in the evolution of brood parasitism in the honeyguides. Because most cuckoos rear their own young and the basal branches in all clades involving brood parasites are birds that build a nest and rear their own young, there is good reason to use phylogeny to estimate the breeding behavior of nesting ancestors in each clade of brood-parasitic cuckoos, and to infer that the nesting behavior of the ancestor was a behavior that facilitated the evolution of brood parasitism. In the cuckoos, brood parasitism has originated independently in the Old World and the New World. New World brood-parasitic cuckoos are more closely related to the terrestrial New World ground cuckoos than to the communal and cooperatively breeding anis and guira, where several

females lay in the same nest and care for the common brood. Old World brood-parasitic cuckoos are closely related to the nesting Old World cuckoos. The independent origin of brood parasitism in Clamator, and its cladistic relationship to the malkohas were unexpected. The basal bird in the Cuculinae is the Raffles’s Malkoha and as far as known it makes its own nest and rears its own young. The closest relatives of Old World brood parasites such as Cuculus are socially monogamous and territorial nesting birds. The molecular phylogeny gives no support to the idea that brood parasites were ancestral to nesting cuckoos, as Hughes (1997a,d) and Aragón et al. (1999) proposed for the nesting Black-billed Cuckoos. Traits that are associated with brood parasitism in the cuckoos include courtship feeding, mimicry in color and pattern of the host egg, and the siblicidal behavior of nestling cuckoos in evicting the host eggs. Courtship feeding that occurs in several lineages of Old World brood-parasitic cuckoos may be retained from courtship feeding in the nesting cuckoos, as this behavior is widespread in nesting cuckoos. Mimicry of the host eggs has evolved at least twice. The New World brood-parasitic Tapera have variable egg colors, and Dromococcyx have spotted eggs, unlike the eggs of nesting cuckoos and somewhat like the eggs of the cuckoos’ hosts. The Old World cuckoos evolved mimicry at least twice, once in Clamator (Clamator jacobinus in part of its range has white eggs; most Clamator eggs are blue like their hosts’ eggs, and unlike the white eggs of the nesting cuckoos to which they are most closely related), and at least one other time with the colored and spotted eggs in the main lineage of brood-parasitic Cuculini). Siblicide with the sharp hooked bill evolved in the New World Tapera. Finally, egg eviction by nestling cuckoos appears to have evolved twice, once in the Thick-billed Cuckoo Pachycoccyx and another time in the Old World Cuckoo lineage before the Cuculus and Chrysococcyx lineages diverged from each other.The four crested cuckoos Clamator, Common Koel, Channel-billed Cuckoo and Long-tailed Cuckoo generally do not evict, whereas Pachycoccyx and the Chrysococcyx and Cuculus lineages of brood-parasitic cuckoos evict host eggs from the nest.

The evolution of brood parasitism in cuckoos 165 In the New World, the birds that are most closely related to Tapera and Dromococcyx are the roadrunners and ground cuckoos.The sister group of these New World cuckoos are the anis and Guira Cuckoo, all cooperative breeders. The phylogeny of New World brood-parasitic cuckoos can be interpreted in two ways: if the most recent common ancestors of the neomorphine cuckoos are taken as the referent, then the second hypothesis is supported and cooperative breeding may have been the starting point in the evolution of brood parasitism in the New World cuckoos. If the sister group of the New World broodparasitic cuckoos is taken as the referent, then the first model (facultative brood parasitism) and the third model (nest takeover) are both supported: in Greater Roadrunner, two females may lay in a single nest and a female may lay in the nest of other kinds of birds. The model of cooperative breeding cannot be rejected from the molecular phylogeny. Cooperative breeding evolved only once in the cuckoos (the crotophagines), so there is no evidence in the tree as a whole to support the notion that cooperative breeding occurred in a common ancestor of the neomorphines and crotophagines, nor that the neomorphines have traits of the cooperative breeders. In both Old World and New World cuckoos, the third model of behavior, the usurpation or claiming of old nests or active nests of other species cannot be rejected, as both the roadrunners Geococcyx and

the cuculine Coccyzus and Coccycua cuckoos from time to time take over the nest of other birds. Nevertheless, the first hypothesis of nesting behavior, that of facultative brood parasitism, appears likely to have been an adaptive precursor of obligate brood parasitism in the cuckoos. Both the New World coccyzine cuckoos (including Black-billed, Yellow-billed, Dwarf and Dark-billed Cuckoos) sometimes lay their eggs in the nests of other conspecifics or in the nests of other kinds of birds, as described in the species accounts. In Dwarf Cuckoo and Dark-billed Cuckoo two females may lay in the same nest, but whether these females mated with the same male or intruded on a nesting pair is unknown. Close relatives of the crested cuckoos Clamator include two Philippine malkohas whose nests are unknown. In Sulawesi a young Yellowbilled Malkoha was fed by another species, a sunbird, and this behavior suggests possible occasional laying and rearing of malkohas in a sunbird nest. What is available from the estimate of conditions that led to brood parasitism are observations of occasional brood parasitism in the groups most closely related to the brood-parasitic cuckoos. The phylogenetic relationships of brood-parasitic cuckoos and nesting cuckoos show that Darwin (1859) had the right idea. So did Friedmann (1933), who first argued that brood parasitism evolved independently in the Old World and New World cuckoos.

This page intentionally left blank

PART II Species accounts

This page intentionally left blank

Crotophaginae Genus Guira Lesson Guira Lesson, 1830, Traité d’Ornithologie, livre 2, p. 149. Large cuckoos of the New World with streaked brown plumage, a crest, and eight rectrices (vs 10 in other cuckoos, except Crotophaga

which also has eight).Type, by monotypy and tautonymy, Cuculus Guira Gmelin. Guira is an application of güirá , a Paraguayan Indian term for bird. One species.

Guira Cuckoo Guira guira (Gmelin, 1788) Cuculus Guira Gmelin, 1788, Systema Naturae 1, pt. 1, p. 414. (Brazil) Other names: Anú branco, White Ani, Piririgüa (Azara). Monotypic.

Description ADULT: Sexes alike, appearance shaggy, head ochre with dark shaft streaks, crest orange-rufous, upper back brown, feathers streaked with a white shaft and light gray edges.Wing coverts dark brown with white edges, wing dark brown with inner vane of wing feathers rufous, outer vane brown with a pale gray edge, giving a frosted appearance.The lower back is white, the rump is buff.Tail blackish with a yellowish base and a broad white band at the tips of T2 to T4, T1 dark brown with a yellowish base: Below pale ochre to whitish, throat and breast streaked brown, under wing whitish, darker on the secondaries; bare facial skin yellow to pea green; iris yellow to orange; bill orange to yellow; legs bluish gray. JUVENILE: Similar to adult, except that the tail has a narrower terminal white band (white of T2 is about 18 mm, white of T3 is 22–28 mm, white of T4 is 20 mm; vs 30, 35–38, and 38–40 mm in the adult); iris light gray; bill black and white, or gray. NESTLING: Skin black and face pale orange, above with long yellowish-white hair-like down, dark

brown on the tips of the primaries and whitish on the tips of rectrices; underparts with short white hair-like down; bill light orange with black ridge along culmen and mandibular rami; iris dark brown; feet black; mouth lining bright pink with raised papilla pads, with a white lateral bar just behind the tip and a white spot behind that, a large circle (open anteriolaterally) on each side of the palate, a ridge of small white papillae at the posterior end of the anterior palate groove, and a fringed white bar perpendicular to the choana pars caudalis; the tongue pink with a black band and white tip (UMMZ photos). SOURCES: AMNH, BMNH, CM, FMNH, MVZ, ROM, UMMZ, USNM, ZMA, ZSM.

Measurements and weights Wing, M (n ⫽ 10) 170–187 (178.5 ⫾ 6.1), F (n ⫽ 8) 161–180 (174.3 ⫾ 5.9); tail, M 214–235 (223.2 ⫾ 6.6), F 198–243 (227.5 ⫾ 13.2); bill, M 27–32 (29.2 ⫾ 1.4), F 36–40 (27.88 ⫾ 1.5); tarsus, M 36–40 (38.0 ⫾ 1.6), F 36–40 (38.0 ⫾ 1.8) (UMMZ). Weight, M (n ⫽ 19) 128.6–168.2 (138.7), F (n ⫽ 8) 113.0–168.6 (144.8) (AMNH, USNM). Wing formula, P7 ⫽ 6 ⬎ 5 ⬎ 8 ⬎ 4 ⬎ 3 ⬎ 9 ⫽ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 36 cm. A long-tailed, brown bird with a shaggy crest, streaked above with a rusty

170 Guira Cuckoo Guira guira crown, streaked below on the breast, the tail with a white base and white tip when seen from below; the wings rufous in flight; the birds often in flocks. Tail jerks up and down, sideways and diagonally.

Voice Noisy, with a great variety of notes, including plaintive whistles “pio..pio..pio..pr..prr..prrr”, yodels, descending phrases “kee-ey, kee-ey, kee-ey, kee-orr, keeorr, cure curecure”; flight call a quiet “yew yew yew yew . . .”; also guttural calls, a high gargled trill, and a “creep” (Azara 1809, Friedmann 1927, Eisentraut 1935, Davis 1940b, Fandiño Mariño 1986, 1989, Hardy et al. 1990, Macedo 1992, Sick 1993).

Range and status South America in Bolivia, Brazil (except in forested regions), Paraguay, Uruguay and Argentina (Hudson 1920, Laubmann 1939, Short 1975, Hilty and Brown 1986, Remsen et al. 1986, Remsen and Traylor 1989, Fjeldså and Krabbe 1990, Narosky and di Giacomo 1993, Sick 1993, Novaes and Cunha Lima 1998, de la Peña and Rumboll 1998, Magalhães 1999). An emaciated vagrant was found in Curaçao (ZMA 28750,Voous 1957); the cuckoo is unknown on mainland northern South America. Resident through their range; in southern part of range more numerous in summer than in winter (Cuello and Gerzenstein 1962). Guira have increased their distributional range in the last 100 to 200 years, appearing in open lands when the forests

have been cleared (Azara 1809, Sclater and Hudson 1889, Sick 1993).

Habitat and general habits Second-growth scrub, drier tree and scrub savanna and scrub woodlands, pampas, pastures, fields, coastal dunes; sea level to over 1200 m (Stotz et al. 1996). They live in drier areas than Smooth-billed Anis Crotophaga ani. They feed on the ground and roost in trees. Common in non-forested areas such as Mato Grosso and the Amazonian savannas, and deforested areas near towns and pastures, they disappear when an area becomes forested, and they are absent in forest areas of Amazonia (Sick 1993). Gregarious, group-living, the members of a group breed communally (Davis 1940b, 1942, Reichholf 1974, Salvador 1981, Cavalcanti et al. 1991, Macedo 1992, 1994, Souza 1995, Stotz et al. 1996). Live and feed in flocks; mean size 6 to 8, with 20 birds in the largest flocks; the members of a pair preen each other (Davis 1940b, Durrell 1956). Birds huddle together on cold days and perch together in a tree at night. On some nights the birds of a flock roost together or in sub-groups. The cuckoos have heavily pigmented black skin between the dorsal feather tracts (UMMZ 236027). On cool days, the wings drop to the side with feathers raised, the black skin exposed to the sun (Sclater and Hudson 1889, Durrell 1956, Storer 1989). Night after night the birds roost in the same trees, huddling together, facing outward, with as many as 14 birds in a clump. This group roosting behavior may help the birds save energy; nevertheless some birds die during cold nights (Friedmann 1927, Davis 1940b, Belton 1984, Gallardo 1984). In spring some groups split into pairs to breed (Sclater and Hudson 1889, Sick 1993). Groups defend territories from other social groups, and some nesting pairs hold breeding territories within these areas and defend these from other pairs in their group (Davis 1940b, 1942, Macedo 1992, Macedo and Bianchi 1997a). While feeding, the birds move from a perch to the ground, or they stalk across the ground in a group; they walk, jump and run after prey which when captured, they swallow whole (Martins and Donatelli 2001).

Guira Cuckoo Guira guira 171

Food Insects, mainly grasshoppers; also cicadas, flying termites, frogs, small reptiles; eggs and nestlings of small birds such as Pied Water-tyrant Fluvicola pica and Fork-tailed Flycatcher Tyrannus savanna, (Azara 1809, Sclater and Hudson 1889, Wetmore 1926, Friedmann 1927, Belton 1984, Gallardo 1984, Mason 1985, de la Peña 1995, Martins and Donatelli 2001). They also take frogs, mice and rats in captivity (Pagel 1992, R. Macedo, in litt.). Nestlings are fed invertebrates and lizards. Most items brought to the young in the nest are ⬍ 6 cm; though some are ⬎ 12 cm (Melo and Macedo 1997).

Displays and breeding behavior The birds live in pairs and in groups throughout the year. In agonistic behavior the guiras raise and open their wings, erect their head feathers in a crest and parade up and down in front of each other with a specific call. One bird preens another, the preening bird with its head feathers sleeked and the tail feathers overlapped into a narrow length, and the head and body held lower than those of its social partner; the bird receiving the preening in a more erect posture, fluffing its head feathers, erecting its crest and closing its eyes, and spreading its tail (Fandiño Mariño 1986, Figure 9.1). Pairing behavior is simple and takes place during the everyday social life of the group without obvious sexual display. Copulation has not been seen and is apparently infrequent. Paired birds carry a leaf in the bill and take the material to a nest site, then bring in more sticks and leaves. Birds are vocal both when nesting and when in nonbreeding groups (Davis 1940b, Gallardo 1984, Macedo 1992; R. Macedo, in litt.). The birds do not appear to behave aggressively in the nest (R. Macedo, in litt.).

Breeding and life cycle In tropical Brazil the cuckoos breed in the dry months May–August, near Brasilia mainly during the rains from August to November (also December to March: Melo and Macedo 1997), near Rio de

Janeiro from August to November, and in Rio Grande do Sul from November to February (Snethlage 1928, Belton 1984). In Uruguay they breed in November and December (Narosky and Yzurieta 1987), and in Argentina from October to February (de la Peña 1995).The nest is a large open platform of sticks and twigs, 15–40 cm in diameter and 14–20 cm high, placed high in a thorny tree or a cactus. Guiras often use and renovate an old nest from the previous season, and add green leaves to line the nest (Friedmann 1927, de la Peña 1995). They sometimes take over old unused nests of other species of birds such as mockingbirds Mimus and add nest material to these (Daguerre 1924). Sometimes guiras lay in nests of other birds, even those of the caracara Milvago chimango and lapwing Vanellus chilensis (Serie 1923a,b, Daguerre 1924, Eisentraut 1935, Sick 1993, Jenny 1997). Eggs are blue to greenish blue and are covered with whitish chalky splotches and streaks raised in relief. Egg size is remarkably variable within a population, 37.2 to 50.8 mm in length, width of 2.78 to 3.86 mm, a mean of 42.7 ⫻ 31.9 mm (n ⫽ 560 eggs); egg weight is 1.67 to 3.22 g (mean 2.46 g, n ⫽ 451 eggs) (R. Macedo, in litt.). Eggs laid by a single female are as different in size and appearance as eggs laid by different females. A few nests have only one breeding pair, whereas in most groups several females lay in the same nest, the number of eggs being larger in the larger groups. Clutch size averages 8–10, ranging from 4 to 30 eggs in a nest. During the period of laying, it is common to find 2, 3 or even 4 eggs laid on the same day (Cariello et al. 2002, 2004).The incubation period is 12–13 days (Macedo and Bianchi 1997a), for 15 days, (de la Peña 1995). The young usually hatch on the same day, or as much as four days apart. The nestling period varies with nesting conditions: the young leave the nest by day 5–6 when disturbed, whereas in undisturbed nests they fledge at day 12–15.The young beg with loud calls and flap their wings when an adult approaches with food. Nestlings that fall from the nest climb up the tree back to the nest, using their spread wings to grasp the trunk, even in an Auracaria where the trunk is covered with spiky leaves. Young guiras

172 Greater Ani Crotophaga major are malodorous in the nest, perhaps due to their carnivorous diet; they leave the nest when more than half grown and are fully adult in size within a month (R. Macedo, in litt.). Parental care continues for three weeks after fledging. Nesting success: 26% of eggs and 55% of hatchlings survive to fledge. Eggs and nestlings are removed or killed by certain adults in the breeding group, and this is the greatest source of loss of eggs and chicks in the nest (Gallardo 1984, Macedo 1992, 1994, Quinn

et al. 1994, Macedo and Bianchi 1997b, Macedo and Melo 1999). The nests are occasionally parasitized by Screaming Cowbirds Molothrus bonariensis (de la Peña 1995). Across one to three nesting attempts within a year, breeding groups lay on average 13 eggs in a season (Cariello et al., in press). Guira Cuckoos have bred in captivity, and hand-reared birds socialize with humans and become inquisitive pets (Sclater and Hudson 1889, Leverkühn 1894, Pagel 1992).

Genus Crotophaga Linnaeus, 1758 Crotophaga Linnaeus, 1758, Systema Naturae (ed. 10) 1, 105. Large cuckoos of the New World with black plumage, a deep and laterally compressed bill, no crest, and eight rectrices (vs 10 in other cuckoos, except Guira which also has eight).Type, by monotypy, Crotophaga ani Linnaeus 1758. All species are

cooperative breeders, either usually or part of the time. The name Crotophaga means “tick eater” (Gr. kroton, tick; phagos to eat) and some are called “tick birds” as they associate with cattle and eat ticks, or did before the period of chemical pesticides.Three species.

Greater Ani Crotophaga major Gmelin, 1788 Crotophaga major Gmelin, 1788, Systema Naturae ed. 10, 1, p. 363. (Cayenne) Monotypic.

NESTLING: Naked, skin black, corners of mouth yellow. The growing black feathers retain a tip of the natal sheath, which is lost at fledging.

Description

SOURCES: AMNH, ANSP, BMNH, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM.

ADULT: Sexes alike, plumage glossy black, the narrow nape feathers and the broad back and breast feathers with a dull blue gloss and edged with a bright glossy bronze-green; wing black glossed blue, tail long, broad, black glossed purple, rectrices broad at the tip; iris greenish white; bill black, compressed with a high arched ridge on upper mandible, with two parallel grooves, one along base of ridge and the other from nostril towards tip of bill; feet black. JUVENILE: Wing glossed blue, tail glossed purple, the rectrices narrower at the tip than in the adult, the bill compressed but without the elevated ridge; iris dark brown.

Measurements and weights Wing, M (n ⫽ 10) 184–217 (202.6 ⫾ 8.7), F (n ⫽ 8) 192–211 (201.0 ⫾ 6.7); tail, M 252–282 (271.8 ⫾ 9.1), F 248–276 (262.6 ⫾ 10.2); bill, M 46–48 (47.2 ⫾ 0.9), F 43–46 (45.1 ⫾ 1.5); bill depth at nostrils, M 22.1–24.7 (23.3 ⫾ 0.8), F 21.0–24.4 (22.1 ⫾ 1.1); tarsus, M 42–44 (42.6 ⫾ 1.0), F 38–41 (39.4 ⫾ 0.9) (UMMZ). Weight, M (n ⫽ 11) 139.5–259 (171.2), F (n ⫽ 3) 145.1–156 (151.6) (AMNH,ANSP, FMNH, MVZ, USNM, UMMZ). Wing formula, P7 ⬎ 6 ⬎ 8 ⫽ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Greater Ani Crotophaga major 173

Field characters Overall length 46 cm. Long-tailed black bird with a crested and compressed bill and pale greenish to white eye (dark in other anis); larger than other anis and a stronger flier. Ridge on bill is highest well anterior to nostrils, and the bill is tapered (rounded in Smooth-billed Ani C. ani). Juveniles with dark eye are identified by their association with adults.

Voice Noisy, a guttural to bubbling, melodious “korokoro” or “toodle-doodle-doodle” given by several birds in a group when they fly or gather on a branch. Flock call, a loud croak “kqua”; alarm call, a harsh raspy note; other calls variously described as growls, croaks, grates, hisses and whirrs “shhhrrrrrrrr” (Davis 1941, Wetmore 1968, Hilty and Brown 1986, Hardy et al. 1990, Sick 1993). In the Orinoco region of Venezuela the sound of bubbling, boiling water gives the birds their local name, hervidor, derived from the Spanish verb hervir, to boil (Cherrie 1916).

1984, Cuello and Gerzenstein 1960, Hayes et al. 1994, de la Peña and Rumboll 1998, Hinkelmann and Fiebig 2001). Seen in São Paulo state, Brazil, in the 1950s, but not in the 1970s or later; locally extinct (Magalhães 1999). Fairly common in much of the range; uncommon south of Tropic of Capricorn, only two records in 100 years in Buenos Aires Province, Argentina (Narosky and di Giacomo 1993).

Range and status

Habitat and general habits

South America, from E Panamá (on the Caribbean slope from western Colón eastward; on the Pacific slope from the Canal area eastward), Colombia, Venezuela, Trinidad, and the Guianas southward through tropical South America east of the Andes to E Ecuador and E Peru, E Bolivia, Paraguay, Brazil, Uruguay and N Argentina. Common along the upper Amazon River in Brazil (Gyldenstolpe 1945a), occasional elsewhere in Amazonia with a record of a lone bird suggesting dispersal (CohnHaft et al. 1997). A population occurs in SW Ecuador at S Los Ríos, Guayas, east of Guayaquil and coastal El Oro (Phelps and Phelps 1958, Wetmore 1968, Davis 1993, Sick 1993,AOU 1998, Ridgely and Greenfield 2001). In addition to the birds in this range, two birds were taken in 1960 at Río Tamesí,Tamaulipas, Mexico, 1200 km north of the nearest known wild population in Panamá (Olson 1978). Resident in most of their range, or with local movements, breeding birds seasonally migratory in Paraguay south to Argentina (Belton

Flooded and riverine habitats, tropical evergreen forests near water, river-edge forests by oxbow lakes, gallery forests, in dense vegetation, thickets and trees along rivers, lakes, swamps and mangroves, wet forest edges, lake edges, grassy edges, thick vine growth and shrubbery, bamboo, marshes, often in vegetation overhanging water. They move into clearings and around houses, sometimes tame around humans; not far into forests. In tropical lowlands they occur mainly below 500 m, casually much higher in the savannas in postbreeding movements or migration, to 2600 m in the eastern Andes of Colombia, and to 2550 m in Cochabamba, Bolivia (Cherrie 1916, Wetmore 1968, Short 1975, Remsen et al. 1986, Remsen and Traylor 1989, Fjeldså and Krabbe 1990,Tostain et al. 1992, Haverschmidt and Mees 1994, Stotz et al. 1996, 1997). Like other anis, the body has a pungent odor. Gregarious, feed in flocks, occasionally follow army ants and squirrel monkey groups which flushout insects. Anis are arboreal and

174 Smooth-billed Ani Crotophaga ani terrestrial ground-gleaning insectivores (Willis 1983, Fjeldså and Krabbe 1990, Sick 1993). Groupliving, two or more pairs live and breed communally in loose colonies.A group defends its territory against other groups of anis (Belcher and Smooker 1936, Davis 1941, Robinson 1997).

Food Insects, mainly orthoptera (grasshoppers, roaches), caterpillars (Brassolidae), beetles, occasionally odonates; other arthropods, spiders, small vertebrates including fish; fruit, berries and euphorbia seeds (Wetmore 1968, Hilty and Brown 1986, Terborgh et al. 1990, Haverschmidt and Mees 1994, Lau et al. 1998).

Breeding In Trinidad they breed from August to November (Belcher and Smooker 1936), in Guyana from May to December (Young 1929), in Suriname from April to September (Haverschmidt and Mees 1994), in French Guiana in April (Tostain et al. 1992) and in Brazil (Pará) in May (Stone 1929). The nest is a bulky mass of sticks and vines, lined with green leaves, built on branches hanging 3–5 m over water. Eggs are greenish blue with a white chalky surface, subspherical or oval, 45 ⫻ 38 mm (Belcher and

Smooker 1936); egg weight 2.4 g (Hellebrekers 1942). Clutch size in Panamá is 2–3 (Wetmore 1968), in Venezuela the average clutch is 7.6 (Lau et al. 1998), in Brazil 3 (Stone 1929) and in Argentina 4–5 (de la Peña 1986). Clutches with more (e.g. 10) eggs in a nest may be eggs of more than one female. Broods with as many as seven young are successful and all young fledge from the nest (Willis and Eisenmann 1979). The incubation period is about 13–14 days.Adults defend nests only after the eggs have hatched; then the ani group successfully drives away Snail Kites Rostrhamus sociabilis and Black-collared Hawks Busarellus nigricollis. Nestlings when disturbed by day 5 or older are able to jump from the nest into the water below, and swim actively on the water surface for several meter to shore, where they run on the ground and conceal themselves in vegetation, then climb back into their nests (Lau et al. 1998). Nestlings also expel large amounts of dense cloacal fluid which has a strong odor and may deter a predator. Nestlings when undisturbed fledge 8–10 days after hatching and remain near the nest for a few more days. Breeding success is greater in nests in isolated sites than in continuous riparian vegetation (av. 4.8 fledged young, vs 1.3 young), due to more eggs hatching in the isolated nests (Lau et al. 1998).

Smooth-billed Ani Crotophaga ani Linnaeus, 1758 Crotophaga Ani Linnaeus, 1758, Systema Naturae (ed. 10), 1, p. 105. [ Jamaica] Other names: Black Witch, Tick Bird, Savanna Blackbird Monotypic.

Description ADULT: Sexes alike, plumage glossy black, head and nape feathers with bronze edges contrasting with the glossy bluish feathers on the back, the feather streaking wide and the tips of the feathers pointed, giving a curved appearance; the nape feathers stiff, wing black with a slight violet gloss, tail long, black, the rectrices broad and truncate; skin around eye black, iris brown to black; bill

arched high and laterally compressed into a thin keel, swollen at base, the arch more pronounced in male than female with the arch half the length of the bill; black, often with pale patches due to frequent scaling, some with a few shallow creases, the lower mandible with an angle on the lower edge; legs black. JUVENILE: Plumage dull black, rectrices more narrow than those of adult and the tips tapered; bill with shallow keel. NESTLING: Naked, skin black; palate pink with white raised patches, one in front and one on each palate fold, row of white marks behind tongue and

Smooth-billed Ani Crotophaga ani 175 U-shaped mark under the tongue; tip of tongue is black, edge of mouth whitish, egg-tooth is present; sheathed pin-feathers grow in a few days (Skutch 1966, Sick 1993, Quinn and Startek-Foote 2000).

calls, with certain calls given in distinct social contexts (Davis 1940a).

SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM.

South Florida, Bahama Islands, West Indies, Caribbean islands; mainland and islands of Quintana Roo in Mexico, Honduras, Nicaragua, Costa Rica, Panamá, Trinidad, Tobago, South America on the west coast south to Ecuador and Peru, and east of the Andes from Colombia, Venezuela, the Guianas and Brazil through the eastern lowlands to northern Uruguay and the Chaco region of Paraguay to northern Argentina (Misiones, Cordoba, Buenos Aires) and the Galapagos Is (Ridgway 1916, Blake 1956, Cuello and Gerzenstein 1962,Wetmore 1968, Short 1975, Remsen et al. 1986, Buden 1987, Benito-Espinal 1990, Fjeldså and Krabbe 1990, Narosky and di Giacomo 1993, Sick 1993, Biancucci 1995, Howell and Webb 1995, CohnHaft et al. 1997, Novaes and Cunha Lima 1998, Raffaele et al. 1998). In Ecuador and Peru they occur both east and west of the Andes, and are replaced in drier areas in the west by the Groovebilled Ani C. sulcirostris (Fjeldså and Krabbe 1990). Resident or locally nomadic. Local groups join together in flocks of 15–35 ⫹ (a record 65) birds in winter, while other groups remain in their breeding territories in all seasons (Loflin 1983). Anis are one of the first birds to colonize newly cleared areas. In Florida, anis became established and bred after a hurricane in 1926, then spread north in the 1960s and bred on the east and west coasts north to Tampa Bay (Merritt 1951, Fisk 1979, Stevenson and Anderson 1994). In the 1970s they became less common and now are local in the Florida Keys and north to West Palm Beach on the east coast and Collier County on the Gulf coast. After the breeding season, some birds disperse out of their breeding range and appear northward and west along the coasts and occasionally inland, to New Jersey, Pennsylvania and Ohio, but records of these are rare (McLean 1995, McLean et al. 1995, Mlodinow and Karlson 1999). Anis appear on the Dry Tortugas, southwest of Florida, apparently arrived from Cuba. In Central America the anis expanded their range during the 20th century.They

Measurements and weights Paraguay: Wing, M (n ⫽ 12) 143–159 (141.8 ⫾ 29.9), F (n ⫽ 12) 136–156 (145.9 ⫾ 6.0); tail, M 172–191 (180.8 ⫾ 5.8), F 160–188 (173.1 ⫾ 8.7); bill, M 30–33 (31.8 ⫾ 1.1), F 28–31 (29.6 ⫾ 1.00); bill depth at nostril, M 20.9–23.4 (21.9 ⫾ 0.9), F 19.2–21.4 (20.4 ⫾ 0.7); bill smaller in juveniles; tarsus, M 32–38 (34.2 ⫾ 1.8), F 32–35 (33.5 ⫾ 0.9) (UMMZ). Weight, Panamá: M (n ⫽ 11) 81.9–133.1 (110.5), F (n ⫽ 11) 77.8–115.6 (92.6) (UMMZ). Wing formula, P7 ⬎ 6 ⬎ 8 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 35 cm. Black birds in flocks; large arched bill, the hump or arch extending above the forehead and in front of a notch between forehead and base of bill, the bill without grooves (sometimes shallow grooves on basal half of bill), the lower mandible gonys with an angle; the head and nape with a bronze gloss, contrasting with the darker green gloss of the back; tail relatively shorter than in the smaller Groove-billed Ani C. sulcirostris. The differences in bill shape of the two species are not apparent in juveniles. Feed on the ground, head and tail held high, outline concave above.

Voice Call, a whining whistle “ani”, “ah-nee”, “ooooeee” or “coo-reek keyoreek”, rising in pitch and given in flight, a querulous “que-lick”; also chatters, mewing with a thin descending “teeew”, growls and clucks (Gosse 1847, Gundlach 1874, Bendire 1895, Hilty and Brown 1986, Stiles and Skutch 1989, Downer and Sutton 1990, Hardy et al. 1990, Sick 1993, de la Peña and Rumboll 1998, Reynard and Sutton 2000).Anis have as many as 13 different

Range and status

176 Smooth-billed Ani Crotophaga ani

first appeared in southern Costa Rica in 1931; by 1975 they were in Guanacaste and replaced Groove-billed Anis in the southern half of the Pacific slope, and they expanded their range in Panamá from the 1940s to 1960s with the human activity of clearing the forests (Wetmore 1968). In the West Indies anis are common in the Bahamas, Greater Antilles, Virgin and Cayman Islands and Providencia; in the Lesser Antilles common on Dominica, St.Vincent and Granada; uncommon on Martinique and Guadeloupe; and rare or absent on other islands. Numbers fluctuate on small islands where birds are absent for a time, then recolonize and disappear again (Terborgh et al. 1978, Raffaele et al. 1998). One population survived a hurricane in 1989 with no decrease in numbers, when anis could feed on weakened migratory prey (Wauer 1989). Recent settlements in the Galapagos may be due to introduction by farmers (Harris 1973, 1982, Rosenberg et al. 1990). Common in much of their range.

Habitat and general habits Second-growth scrub, river island scrub, humid areas, clearings in forests, scrub and scattered trees in pastures, woodland, thickets, mangroves, lake margins, grassy edges, roadsides, marshes, thick vine growth, Cecropia and dying trees, and marshes, most common where they occur in outskirts of towns and old cultivation. In Puerto Rico they live in disturbed dry habitats where the vegetation is dominated by exotic grasses (buffel grass Centris ciliaris, guinea grass Panicum maximum) with scattered trees, often mesquite.They feed in fields of grass and use

their high crowned bill to separate the wet, dense vegetation as they forage (Willis 1983b, Quinn and Startek-Foote 2000). They live mainly in lowlands, from sea level to 500 m, occasionally higher (Stotz et al. 1996), to 2400 m north of Orinoco in Venezuela, to 2000 m or more in Colombia, 2400 m in southern Ecuador, and breed in the temparate side valleys of the Marañón Valley in northern Peru (Fjeldså and Krabbe 1990). In South America, their heart and lungs are longer, and body weight greater when they are above 2000 m in the Andes, than in the lowlands (Köster 1976). Anis live in small, noisy flocks of 2–11, occasionally of as many as 30 birds, each bird a few meters from the others (Loflin 1983) as they walk or hop on the ground and on branches, and rise into the air and fly in pursuit of insects. On the ground they feed with cattle, moving among the feet and head of the cattle. They take ticks from the legs of cattle and they sometimes perch on cattle (Gosse 1847,Wetmore 1916, 1968). Anis follow cattle, horses and indigenous mammals (tapu Dasypus sabanicola, de Visscher and Moratorio 1984); they follow a plow, and also army ants (Willis 1983b). They capture insects that are flushed out; and they catch other small animals that flee from grass fires. A bird finds food at twice the rate when it follows cattle, than when it feeds alone (Smith 1971). The flight is straight and rapid with shallow wingbeats and a short glide. Large groups of anis have larger territories than small groups. In Puerto Rico, a 3-adult group had territory of 6.2 ha, a group of 5 adults had 8.7 ha, and a group with 6 adults (increasing to 9) had 9.6 ha (Quinn and Startek-Foote 2000). Once established, a group can hold its territory against challenges from other groups, sometimes even when the resident group size changes (Davis 1940a). Anis often spread their wings and tail to sun after they wet their plumage while climbing through wet places and dense herbs; they appear lean and shabby (Bendire 1895).They allopreen, touch each other, expose their skin to the sun while they spread their wings and fluff their feathers. They huddle together during the rain, and at night they roost in marshes, huddling to save heat; at night their body temperatures become 8°C cooler than

Smooth-billed Ani Crotophaga ani 177 in the daytime; and they bask in the sun, with wings spread, to warm their bodies in the morning before they become active (Bent 1940, Merritt 1951,Warren 1960,Wetmore 1968).

Food Insects, mainly grasshoppers; also mantids, other orthoptera including mole cricket nymphs, cockroaches, beetles including fireflies and weevil rootborers which are pests of sugar cane and citrus orchards, squash bugs and assassin bugs, dragonflies, moths and caterpillars (Brassolidae, Noctuidae, Geometridae), butterflies (Pieridae), hymenoptera (euglossine bees), but few ants and spiders (Gosse 1847, Gundlach 1874, Wetmore 1916, Danforth 1925, Friedmann 1927, Stockton de Dod 1981, de Visscher and Moratorio 1984, Sick 1993, Haverschmidt and Mees 1994, Rosenberg et al. 1990, Magalhães 1999, Burger and Gochfeld 2001). Recent observations show that few ticks are taken (Wetmore 1916, 1968, Rand 1953), but in the 1800s before chemical pesticides were used on livestock, ticks were important in their diet (Gosse 1847). Anis plow into soft earth and cow dung to get insects, and they chase flying insects.They take tree snails, Anolis lizards, small snakes, frogs, mice, nestling birds and eggs (anis are mobbed by other birds: Wetmore 1916, 1927, Danforth 1925, Dubs 1982, Loflin 1983, de Visscher and Moratorio 1984); and they take small birds from mist-nets (Gill and Stokes 1971). In the dry season they take fruits, including snake-withe berries, fiddle-wood Cytharaxylon and wild grape Cissus sicyoides (Gosse 1847, Quinn and Startek-Foote 2000), Madame Jeanette hot peppers (Haverschmidt and Mees 1994) and seeds (Davis 1940a, Downer and Sutton 1990, Sick 1993).They eat the fruits of the gumbolimbo Bursera simaruba, removing the skin, swallowing the inner part and regurgitating the seed (Trainer and Will 1984).

Displays and breeding behavior Breeding adults are often socially monogamous; rarely polygynous, a male mating with two females, or polyandrous, a female copulating with two males when the earlier mate disappears. Anis copulate

both in the nest tree and in other trees. In mating, the male simply mounts the female. In copulating, the female remains stationary and raises her tail, then pulls her wings together in front while the male flaps his wings. Copulation lasts as long as 42 seconds. A pair copulates with either sex on top, males mounting females as well as females mounting males. Copulation occurs throughout the breeding cycle but is most common in the nestbuilding period. In courtship feeding, the male presents the female with a grasshopper or lizard, copulates and then she swallows the food, or the female accepts the food without copulating. Besides feeding the female during her fertile period, the male feeds the female while she incubates (Köster 1971, Loflin 1983, Quinn and Startek-Foote 2000). Social hierarchies are apparent in the nest, where one bird displaces another by pushing her bill underneath the sitting bird, and this has the effect of delaying the onset of incubation of the clutch until the last female has laid her eggs (J. S. Quinn, in litt.).

Breeding and life style In Florida they nest after the rains begin in May through October. Where rains are not seasonal in the West Indies, Suriname and French Guiana, the anis may breed all year round. In more seasonal areas of the Neotropics they breed during the rains, in SW Puerto Rico starting in September after the rains have begun, in Cuba from April to October, in Central America during the rains after the herbs have grown and grasshoppers are abundant, and in southern Brazil in January and February (Gundlach 1874, Wetmore 1916, Danforth 1925, Belcher and Smooker 1936, Hellebrekers 1942, Loflin 1983, Belton 1984, Tostain et al. 1992, Haverschmidt and Mees 1994, Raffaele et al. 1998). On Grand Cayman Island, where the wet season occurs from May to November, they breed throughout the year (Bradley 2000). Gregarious, group living with several pairs breeding together in a single nest.The nest is built and attended by a single pair or by 2–4 pairs with several females laying in a common nest. As many as nine adults may take part in building the nest. In Florida, anis nest either in pairs or in groups,

178 Groove-billed Ani Crotophaga sulcirostris whereas in other tropical areas, anis are communal and competitive (Gosse 1847, Eisentraut 1935, Bent 1940, Davis 1940a, 1942, Skutch 1966, Köster 1971, Reichholf 1974, Loflin 1983, Stiles and Skutch 1989, Sick 1993, Stevenson and Anderson 1994, Quinn and Startek-Foote 2000). The nest is built in a thorny tree, shrub or thicket, a large, bulky shallow mass of interlaced sticks. It is 17–26 cm in outside diameter, 13–19 cm inside, the inside depth being 5–9 cm (Quinn and Startek-Foote 2000). Green leaves are added each day before laying, during laying and during incubation. Eggs are greenish blue covered by a white chalky surface, 35 ⫻ 26 mm in size; weight 1.15 g (Hellebrekers 1942), and are quite variable within a population (34.5 to 37.2 ⫻ 26.0–28.6 mm; weight 10.8 to 13.0 g) (Loflin 1983). Eggs are laid at 2-day intervals. Clutch size averages 9.7, varying with the number of females at the nest, where nests with one laying female average 4.3 (range, 3–7), nests with two females 9.5 eggs, nests with 3 females 14.8 eggs, and with 4 females 21.3 eggs; as many as 36 eggs are laid in a nest. In nests where more than one female lays, the female covers with leaves the eggs of the female that laid earlier, before she lays her own.When the early females attempt to raise these eggs to the top layer, the later-laying females chase the early females from the nest and continue to bury the early eggs. As a result, often only the top layer of eggs hatches. Occasionally a female tosses eggs of an earlier female from the nest. In communal nests,

37% of eggs are lost when buried in the nest lining by competing females. Other sources of loss are infertility, predators and storms. Incubation begins soon after the first egg in single pair nests, but later in multiple-female nests. Competitive behavior occasionally leads to one adult killing another. Both sexes incubate. The incubation period is 13–14 days. Nestlings are likely to starve if latehatched—these may be the young of the last-laying female.The last-laying pair sometimes abandons the nest, and their young are reared by the early-laying pairs that rear their own brood as well.The nestling begs by stretching its head and neck and flapping its wings ( J. S. Quinn, in litt.). Nestling feathers begin to emerge from sheaths in five days.The young can grasp with their feet and crawl when they are disturbed; they scramble out of the nest and cling to branches as early as day 4 ( J. S. Quinn). Nestlings normally fledge at 10–17 days. Both sexes care for the young, as do juveniles that remain in their natal territory for as long as a year. As many as nine anis bring food to the nestlings (Davis 1940a), and both the adults and nestlings defecate over the side of the nest (Quinn and Startek-Foote 2000). Breeding adults give less care to broods with helpers, which give care to the brood. Group size has no effect on feeding rates or growth of the young. The young fledge at 35 g, less than half the weight of adults, and they complete most of their growth after they leave the nest. Most anis disperse from their natal group at 8 to 12 months of age (Oniki and Ricklefs 1981, Loflin 1982, 1983).

Groove-billed Ani Crotophaga sulcirostris Swainson, 1827 Crotophaga sulcirostris Swainson, 1827, Philosophical Magazine (n.s.) 1, p. 440. (Temascaltepec, Mexico) Monotypic.

Description ADULT: Sexes alike, plumage glossy black with a blue gloss due to the iridescent edge of the feathers on the head, nape and back, the feathers narrow and straight with the tips pointed, the skin under

the feathers also black, wing black, tail long, black; skin around eye black, iris brown to black; bill arched and laterally compressed, no hump at the base, the upper ridge continuous with crown, upper mandible with parallel grooves along the bill, extending to near the tip in older adults, the lower mandible nearly straight in profile; feet black. JUVENILE: Plumage dull black, soft and loosewebbed; cheeks bare at fledging, feathers develop

Groove-billed Ani Crotophaga sulcirostris 179 later, tail feathers narrower and more pointed (not truncate) than in adult; bill smaller than adult, without conspicuous grooves, but grooves are present.

hollow “pep, pep . . .” (Skutch 1959, Wetmore 1968, Vehrencamp et al. 1986, Stiles and Skutch 1989, Hardy et al. 1990).

NESTLING: Naked, skin black, no hair-like down when they hatch, but covered with dark downy contour feathers when they fledge; bill with grooves in the older nestlings; palate bright pink with white raised patches, one on each side of the palate fold or choana, white edge on underside of tongue and white line to tip of under tongue, tip of tongue black, edge of mouth whitish.

Range and status

SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, RMNH, ROM, UMMZ, USNM.

Measurements and weights Texas: Wing, M (n ⫽ 12) 129–138 (134.7 ⫾ 3.5), F (n ⫽ 10) 127–135 (131.2 ⫾ 2.7); tail, M 168–184 (177.0 ⫾ 6.9), F 166–177 (171.9 ⫾ 4.0); bill, M 26–30 (29.0 ⫾ 1.3), F 26–29 (27.5 ⫾ 0.8); bill depth at nostrils, M 17.3–19.8 (18.9 ⫾ 0.9), F 16.6–19.1 (17.5 ⫾ 0.9), smaller in juveniles; tarsus, M 31–35 (33.2 ⫾ 1.4), F 31–34 (33.1 ⫾ 1.1) (UMMZ). Weight, Mexico and Panamá, M (n ⫽ 14) 61.1– 94.5 (79.6), F (n ⫽ 22) 55.5–79.8 (66.6) (UMMZ); Costa Rica, M (79), F (70) (Vehrencamp 1978). Wing formula, P6 ⬎ 7 ⬎ 5 ⬎ 8 ⬎ 9 ⫽ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 32 cm. Black birds in loose flocks; bill large and with grooves, arch not extending above the forehead in profile and the lower mandible rather straight; the blue gloss of head and nape uniform with the gloss of the back; the glossy feathers narrower than in Smooth-billed Ani C. ani, and the tail long, often pointed upwards when the birds are on the ground.

Voice Squeaky and chattering high-pitch sounds, “tee-ho tee-ho” in series, the first note slurring up, given in excitement or alarm; also a faster series, a whining, querulous “whee-o”, a long series of rapid whistled “kiw” slurred down at the end, a grating “krr krr . . .”, and a “conk” or sharp,

Gulf of Mexico from central and southern Texas and (once breeding) Louisiana, Mexico (southern Sonora south along both the Pacific and Caribbean slopes and near-coastal islands including Cozumel I and Ambergris Cay off eastern Yucatán Peninsula), south through Caribbean and Pacific Central America and western South America along the coastal lowlands to Ecuador, Peru, northern Chile and NW Argentina. In the southern Caribbean region they occur in northern South America from Colombia (mainly N Córdoba east to Guarija, south in drier parts of Magdalena Valley to N Huila) to Venezuela and Guyana, and on the Netherlands Antilles and Trinidad (Phelps and Phelps 1958, Snyder 1966, Meyer de Schauensee and Phelps 1978, Voous 1983, Araya Mödinger and Mille Holman 1986, de la Peña 1986, Hilty and Brown 1986, Stiles and Skutch 1989, AOU 1998, Ridgely and Greenfield 2001). Nonbreeding movements occur when birds disperse northward into continental North America, as far as panhandle and peninsular Florida, the southwestern states, California, the southern Great Plains and the Great Lakes region to southern Ontario and Manitoba, and an ani is accidental in the Revillagigedo Islands (Socorro) ( Johnson 1967, Sutton 1967, Oberholser 1974, AOU 1998, Kent 1988, Russell and Monson 1998, Mlodinow and Karlson 1999). Although anis are resident in most of their range, after the breeding season some appear as far as 1000 km from their nearest breeding areas. In central United States they appear in autumn from late September to November in Florida and Alabama they appear from September through April and May, and in the southwestern states they appear in all seasons, mainly September to November. In the nonbreeding season they live in flocks and roost in groups of as many as 30–40 birds. Many birds remain on their breeding territories and live in the same groups throughout the year, particularly in the better habitats where insects occur throughout the year; while birds in

180 Groove-billed Ani Crotophaga sulcirostris

xeric habitats such as pasturelands move to riparian areas. A few birds occur in the northern part of their range in winter in northern Mexico, the Gulf Coast of Texas and further east (Root 1988, Bowen 2002). Groove-billed Anis are common in much of their range, but are replaced by Smooth-billed Anis C. ani in the Pacific areas of Central and South America and in the West Indies.Territories of social groups occupy an area of 1–10 ha; birds in marshy areas have larger breeding groups and smaller territories than birds in pastures. Anis formerly lived on the cape and islands of southern Baja California (Grinnell 1928); the last breeding record was in 1896, and since about 1910 only one has been seen there, a vagrant (Howell and Webb 1995, AOU 1998). A subspecies Crotophaga sulcirostris pallidula described by Bangs and Penard (1921) as paler was recognized for birds on the west coast of Mexico in Sinaloa, Colima and Tehuantepec (Bangs 1930), but these birds are not distinct. Even the anis taken at Cape Baja in 1887 do not differ consistently from anis in other regions: some old specimens are brownish due to post-mortem change (van Rossem 1938, MVZ). On the other hand, historically the range has extended to other areas, with birds colonizing Curaçao then Bonaire and Aruba within the past century (Voous 1957).

to tropical scrub, less often in humid regions. Anis occur mainly in lowlands, and they range mainly from sea level to 500 m, occasionally in Ecuador and Peru to 2700 m (Lowery and Dalquest 1951, Wetmore 1968, Hilty and Brown 1986, Fjeldså and Krabbe 1990, Stiles and Skutch 1989, Stotz et al. 1996, AOU 1998, Ridgely and Greenfield 2001). Gregarious, group-living, weak fliers.They feed in groups, walking on dry ground, often in pasture with cattle or other livestock, where they catch insects more rapidly when in groups than when hunting alone (Dickey and Van Rossem 1938, Rand 1953, Skutch 1959, 1983, Smith 1971, Stiles and Skutch 1989). They take ticks from the legs and undersides of cattle, or did so in the days before the widespread use of cattle dips with pesticides. They dig in dung for grubs, dung-beetles and other insects (Voous 1957, 1983).They follow army ants into the forest and capture insects that are flushed out by the ants (Bent 1940, Willis 1983). Territory size of a group is large, 3–11 ha. The local density of birds is higher in marsh than in pasture (1.9 vs 0.4 birds / ha), even though the density of insect prey is about the same in these habitats. As many as four pairs live in a group and lay in a common nest. The number of birds in a group does not affect their territory size or the number of young that survive and fledge in their common nest. However, the number of birds that help rear the young may reduce the costs of parental care to the breeding adults (Vehrencamp 1978, Bowen et al. 1989).

Food Insects, mainly grasshoppers, also cockroaches, grubs, cicadas, flies, wasps, ants, termites; ticks, spiders, small vertebrates including Anolis lizards, and fruits, seeds and berries. Most foods are larger than 4 mm (Bent 1940, Dickey and Van Rossem 1938, Voous 1957, Wetmore 1968, Oberholser 1974, Vehrencamp 1978). The same foods are fed to the young anis in the nest (Bowen 2002).

Habitat and general habits

Displays and breeding behavior

Tropical lowlands in scrub, thickets, pastures, fields, marshes, cleared and disturbed areas in the range of original vegetation from tropical evergreen forest

A contact species like the other anis, the birds roost and huddle together in all kinds of weather. They preen each other: one bird stretches its neck

Groove-billed Ani Crotophaga sulcirostris 181 upward and the other bills and nibbles at the feathers, then the birds switch roles as the preener and the preened. In the early morning they spread their wings and sunbathe, a behavior that accounts for their name “zopilotillo,” a diminutive of “zolpilote” or vulture (Skutch 1959,Wetmore 1968). Agonistic displays are aggressive in nature: (1) Broadside threat, where the threatening bird lowers its head, presents its large bill in profile, drops its near wing and raises its far wing to enlarge its body profile, sometimes erecting its plumage as well, and gives a “conk” call. Males threaten another male that approaches their mate; females threaten other females when they interact at a nest; and adults threaten juveniles of the same sex during the dry season between breeding seasons when the young disperse from their natal territory. The response of the threatened bird is to sleek its plumage, lower its head, and turn or move away, or to solicit preening from the other bird. (2) Aerial chase, which is directed towards potential immigrants and territory invaders from outside the social group, and to rival males within the social group. (3) Supplanting behavior, which occurs near a nest and controls the access of other females to the nest. Anis in a group do not live in a strict dominance hierarchy, and a female’s status changes when alone and when her mate is nearby (Vehrencamp et al. 1986). Anis join a social group in different ways depending on the number of applicants to the group and the number of birds already in the resident social group. (1) A single bird attempts to join a group; it gives a call given only in that context, sometimes persisting in calling for days, while it is threatened and chased away by the dominant resident birds. When a bird repeatedly fails to enter a group, it moves to another group and tries again. A resident group admits a new bird when they have lost a member of the same sex. (2) A pair joins a group with less apparent control over the pair by the resident group; the new pair sometimes enters by laying in the nest of the resident pair, then remains and takes part in rearing the young. (3) A resident pair splits and each bird admits an immigrant of the opposite sex as its new mate (Vehrencamp et al. 1986). Anis breed as early as their yearling year (Vehrencamp 1978).

A nestbuilding pair is socially monogamous.The male and female spend time together, perch in contact and allopreen, call back and forth near the nest, and call over a long distance when one mate leaves the other at the nest. The pair remains together from year to year as long as both birds are in the social group (Vehrencamp et al. 1986). In colormarked birds, nearly all copulations took place between social mates. The few extra-pair copulations were between nesting attempts or were at the end of laying and were unlikely to lead to fertilization. During the breeding season and shortly before the eggs are laid, the male mounts the female. Males develop small testes while they incubate. Reversed mounting with the female on the male sometimes follows the loss of a nest. The female also mounts the male outside the breeding period; during this dry period all the matings observed of individually color-marked anis have been female-on-male (Bowen et al. 1991,Vehrencamp 1982b). In courtship feeding, one bird approaches the mate with an insect in its bill, and it holds its breast up and out. The bird approached lowers its body into a crouch. As the first mounts, the second spreads its wings; the bird on top flutters its wings and wags its tail; then the upper bird’s tail dips and cloacal contact takes place. Females take food offered by the male, but males do not take food when the courtship is initiated by the female (Bowen et al. 1991).

Breeding In Texas from May to August with young in the nest as late as September (Oberholser 1974), in the Cape region of Baja California in August and September (Brewster 1902), in Oaxaca from May to July (Rowley 1984, Binford 1989), in the Caribbean lowlands during the rains from June to early September; in Curaçao the same birds have apparently bred four times in succession from December to September (Voous 1983); in Costa Rica the birds breed in the rainy season of June to November, later in years when the rains are late (Vehrencamp et al. 1986, 1988, Stiles and Skutch 1989), in Colombia they breed in October and January (Hilty and Brown 1986), in Venezuela they breed in July and September to November (Thomas 1979).

182 Groove-billed Ani Crotophaga sulcirostris The nest is built in a marsh, thicket, shrub or tree, often a spiny citrus tree, 3–5 m above ground. It is a large, open bulky platform, measuring about 25 mm across and 10 mm deep, with an inner cup about 10 mm across and 6 mm deep, built of sticks, roots and thorns, and bits of vine and herbs, lined with green leaves which they continue to add during laying and incubation. Occasionally anis take over an old nest of another bird, such as the Boat-tailed Grackle Quiscalus mexicanus, and add a few sticks and leaves (Skutch 1959). A nest is built and by solitary pairs or by 2–4 cooperating pairs. Among these communal breeders, several females lay in a common nest, each female attended by her mate. Eggs are greenish blue covered with a white chalky surface, easily scratched; 31 ⫻ 24 mm (Bendire 1895, Wetmore 1968); 11 g (an egg is c. 17% of female body weight,Vehrencamp 1978). Eggs are laid at any time of day, often in the early afternoon. Clutch: a female lays 3–4 eggs at intervals of 2–3 days. More than one female may lay in a nest, which may have as many as 18 eggs. A female that has not yet begun to lay often tosses the eggs of other females out of the nest.When she has begun to lay she does not remove the other eggs in the nest, suggesting that females do not recognize their own eggs in a nest where more than one female has laid; rather, they change their behavior and avoid removing their own eggs.When a female from outside the social group lays, the group often abandons the nest.The incubation period is 12–14 days. An adult incubates for 30 min to 1 h at a time during the day; both male and female take about the same time. In solitary pairs the male incubates and later broods the young through the night. In communal nesting groups, one male incubates and broods the young at night. He is generally the oldest male in the group, the largest male, and the mate of the last-laying female, which differs from other females in the group in size (longer wing) and age (older) (Skutch 1959, Vehrencamp 1978, Vehrencamp et al. 1986). Nestlings develop rapidly and their eyes open two days after hatching. They are brooded for the first week; then their feathers break through the sheaths and the young look like feathered birds.

Both parents bring food to the brood. Together parents and helpers bring food at a rate of about six feeds per nestling in an hour.Two adults sometimes divide food between themselves at the nest, then each feeds the young. The nestlings beg in an upright posture when they are hungry, with neck stretched upward and wings stretched out and fluttered at the side (B. S. Bowen, in litt.). Nestling anis remain in the nest for 8–10 days, but in asynchronous broods the later-hatched birds are smaller and sometimes die of starvation. Parents remove nestling feces when the nestlings are young, and after the nestlings are two or three days old the young birds squirt their excrement over the side of the nest, which becomes marked with the droppings. Nestling feathers begin to emerge from sheaths in five days, when the young can grasp with their feet and crawl when they are disturbed (B. S. Bowen). Wing length develops independently of the number of young in the brood, or the number of adults in the breeding group (Vehrencamp 1978). After the young leave the nest, they can make short flights from branch to branch by day 11, but often spend 10 to 15 days near the ground. They normally fly by about day 17–20 and are tended by their parents until they are about 6 weeks of age.Young anis remain in their natal area for several months, and when they are about 10 weeks old they sometimes aid in incubation and feeding a later brood in their natal group (Miller 1932, Dickey and Van Rossem 1938, Skutch 1959, Vehrencamp 1977, 1978, 1983, Koford et al. 1986, 1990, Vehrencamp et al. 1986, 1988, Bowen et al. 1989, Stiles and Skutch 1989, Bowen 2002). Young anis that were marked in their natal site and survived until the end of the breeding season often remained in their natal area. Early in the next breeding season, 48 of 196 independent young were still in their natal area; then 25 dispersed within the study area and bred 1–3 territories away from their natal group, while another 15 bred in the group where they were born. Males are more likely than females to remain and breed in their natal group, and more males than females that breed in a group were born there. When a bird leaves the group, it disperses alone rather than with a nestmate (Bowen et al. 1989, Koford et al. 1990).

American Striped Cuckoo Tapera naevia 183 Survival of adults is 66–72% from one year to the next, although some local loss of adults may be due to dispersal rather than to mortality. Most known deaths occur during the breeding season when adults

are at risk on the nest.Anis are taken through the year at night roosts by their predators, Trimorphodon lyre snakes and Vampyrum carnivorous bats (Vehrencamp et al. 1977,Vehrencamp 1978, Bowen et al. 1989).

Neomorphinae Genus Tapera Thunberg, 1819 Tapera Thunberg, 1819, Handlingar Kungl Vetenskaps och vitterheetssamhället i Göteborg , 3, p. 1. New World Cuckoos with a streaked brown plumage, a crest, a long and broad alula, and a long tail with 10 narrow rectrices. Type, by monotypy, Tapera brasiliensis

Thunberg ⫽ Cuculus naevius Linnaeus (Peters 1940). Tapera is a Tupí (Brazilian) name matim tapirera for a cuckoo whose cries represent the voices of ghosts. One species.

American Striped Cuckoo Tapera naevia (Linnaeus, 1776) Cuculus naevius Linnaeus, 1766, Systema Naturae (ed. 12) 1, p. 170. Based on “Le Coucou tacheté de Cayenne” Brisson, Ornithologie 4: 127, pl. 9, fig. 1. (in Cayania ⫽ Cayenne) Other common names: Striped Cuckoo, Cuatro Alas, Quatro Asas,Tres Pesos, Four Winged Cuckoo Monotypic.

Description ADULT: Sexes alike, upperparts brown streaked buff and black, head with a striped blackish and rufous shaggy crest; the alula large and black, wing brown to rufous, white stripe across base of primaries seen in flight; upper tail coverts light rufous becoming dark brown with the loss of the pale feather edges; tail long and graduated; in fresh plumage the brown central vane with light rufous edges gives a sandy appearance but in worn plumage the loose-webbed edges are gone and the tail is dark brown; face with a white eyebrow and black whisker line; underparts whitish with narrow black malar streak, throat and breast with distinct black streaks, belly white; bare skin around

eye yellow, iris brown to greenish; bill brown to orange-brown or yellowish below; feet gray brown. Underparts fade with the season; fresh plumage in autumn is ochraceous, plumage in spring less so, and by midsummer the breast and flanks are a dirty gray. JUVENILE: Crown black with buff spots (not streaks), upperparts rufous brown streaked buff and black with buff spots, alula black with buff terminal spots, upper wing coverts brown with buff spots, rump and upper tail coverts brown with buff spots, tail brown with rufous edge and tips, underparts buff with fine black bars on throat and breast, belly white with small dark spots (not streaks). NESTLING: At hatching naked, skin pinkish; gape orange-yellow, bill with a sharp curved tip; within a week the skin turns darkish violet (Haverschmidt 1961). SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM.

184 American Striped Cuckoo Tapera naevia

Geographic variation Regional populations have been described as subspecies on the basis of size. Nevertheless, measurements of museum specimens show a considerable overlap in size between these regions. Tapera naevia excellens (Sclater, 1858) was recognized in Mexico and Central America to E Panamá, described as smaller in the bill and wing than birds in northern South America, but the size of birds in those regions overlap considerably and the form is not recognized here. A third form, Tapera naevia chochi (Vieillot, 1817), was recognized in southern Brazil south from Mato Grosso and São Paulo and in Paraguay and Argentina, and was said to be larger and different in plumage color (Bangs and Penard 1918, Cory 1919, Pinto 1947), but size and plumage color of birds in this region overlap extensively with those of birds in northern South America (Wetmore 1926, Hellmayr 1929, Traylor 1958, Sick 1993, and measurements) and the form is not recognized here.

Measurements and weights Mexico to Costa Rica: Wing, M (n ⫽ 13) 116–123 (118.8 ⫾ 3.2), F (n ⫽ 8) 107–121 (113.4 ⫾ 4.0); tail, M 171–190 (178.5 ⫾ 8.9), F 152–194 (169.9 ⫾ 13.1); bill, M 19.5–21.7 (20.4 ⫾ 0.6), F 17.1–20.5 (19.1 ⫾ 1.1); tarsus, M 30.8–35.9 (33.4 ⫾ 1.7), F 30.0–34.1 (31.8 ⫾ 1.5) (AMNH, UMMZ); Panamá: Wing, M (n ⫽ 10) 108.1–117.5 (112.4), F (n ⫽ 10) 104.0–112.3 (108.2); tail, M 148.0–165.0 (157.7), F 140.0–162.0 (146.2); bill, M 21.0–23.2 (21.8), F 20.2–22.6 (20.8); tarsus, M 32.3–36.1 (34.2), F 30.5–33.0 (32.0) (Wetmore 1968); Colombia, Venezuela, the Guineas, N Peru, Ecuador and Brazil to River Amazon: Wing, M (n ⫽ 13) 103–125 (110.3 ⫾ 6.4), F (n ⫽ 9) 101–109 (105.1 ⫾ 3.2); tail, M 143–169 (154.8 ⫾ 9.5), F 136–152 (148.8 ⫾ 7.8); bill, M 16.4–19.7 (18.1 ⫾ 0.9), F 15.3–18.7 (17.0 ⫾ 1.1); tarsus, M 25.5–34.4 (30.7 ⫾ 2.3), F 26.9–32.0 (29.5 ⫾ 1.9) (AMNH, UMMZ); Paraguay: Wing, M (n ⫽ 11) 110–118 (113.9 ⫾ 3.3), F (n ⫽ 4) 106–110 (108.3 ⫾ 1.7); tail, M 153– 173 (163.4 ⫾ 6.2), F 142–146 (144.0 ⫾ 2.3); bill, M 17.7–19.8 (18.7 ⫾ 1.0), F 15.2–17.4 (16.5 ⫾ 1.0);

tarsus, M 26.0–33.0 (30.3 ⫾ 2.3), F 26.7–30.9 (28.8 ⫾ 1.7) (AMNH, UMMZ). Weight, M (n ⫽ 8) 43–56.5 (49.4), F (n ⫽ 7) 41–59 (51.1) (AMNH, MVZ, UMMZ); Bolivia, M (n ⫽ 2) 40.0–47.7 (Schmidt et al. 1997). Wing formula, P7 ⬎ 6 ⬎ 8 ⫽ 5 ⬎ 4 ⬎ 9 ⫽ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 26–29 cm. Long-tailed cuckoo with streaks, brown back striped black and a blackstreaked rufous crest. In flight the wing is white across the base of the primaries.When perched the bird is erect, raises and lowers its crest and swings its tail from side to side.

Voice The most common song is a mellow whistle, the second note half a tone higher than the first, often repeated,“sem-fim” or “fee-fee”, repeated or sometimes with a third note, each note at nearly 3 kHz and lasting 0.5–0.6 sec, and the song lasting nearly 1.5 sec. A second song has an ascending series of 4–7 short whistled notes,“tres pesos pide” or “peepee-pee-PEEdee”. The song rises through the first five notes from 2.6 to 3.0 kHz, the last note “dee” drops to 2.7 kHz; the song lasts about 2 sec. A third song is a soft variant of the second, with fewer notes (2–4), and other soft sounds are given. Songs in Central America, Colombia, Venezuela and southern South America are similar (Friedmann 1927, Slud 1964, Belton 1984, Hilty and Brown 1986, Stiles and Skutch 1989, Hardy et al. 1990, de la Peña and Rumboll 1998, Smith and Smith 2000). Regional differences in song as reported in Brazil (Sick 1953b, 1993) may be behavior variants as are known within an area and in a single bird (Hardy et al. 1990, Haverschmidt and Mees 1994, Smith and Smith 2000). The young cuckoo gives a short whistle like that of its host young (Morton and Farabaugh 1979).

Range and status Southern Mexico (Veracruz, Oaxaca, Tabasco, Chiapas and Quintana Roo), south through Central America along the Caribbean and Pacific slopes to eastern Panamá (AOU 1998), and in

American Striped Cuckoo Tapera naevia 185

South America from Colombia,Venezuela and the Guianas (Guyana, Suriname and French Guinea), western Ecuador (Esmeraldas south through El Oro and W Loja), northern Peru, and east of the Andes in Brazil (lower Rio Madera, Goais, Bahia, Mato Grosso and São Paulo) south to Bolivia, Paraguay, Uruguay and northern Argentina (Tucumán, Buenos Aires); and on Margarita and Trinidad (Chubb 1916, Wetmore 1926, Hellmayr 1929, Pinto 1938, Lowery and Dalquest 1951, Cuello and Gerzenstein 1962, Short 1975, 1976, Meyer de Schauensee and Phelps 1978, Hilty and Brown 1986, Remsen and Traylor 1989, ffrench 1991, Tostain et al. 1992, Narosky and di Giacomo 1993, Sick 1993, Haverschmidt and Mees 1994, Novaes and Cunha Lima 1998). Resident; in southern part of range their numbers vary with the season (Cuello and Gerzenstein 1962, de la Peña and Rumboll 1998, Magalhães 1999). The species range has expanded in the past 50 years as birds moved into cleared and scrub lands in response to deforestation in Central America and in Brazil (Mitchell 1957, Belton 1984, Stiles and Skutch 1989, Skutch 1999) and have appeared at high altitudes and into southeastern Ecuador in cleared habitat (Ridgely and Greenfield 2001). Common to uncommon, shy and solitary.

Habitat and general habits Shrubby clearings, scrub, low seasonally wet grassland, overgrown pastures, mangroves, river island scrub, open country with scattered trees, thickets and bushes, clearings in forested areas, and brush at

edge of forests.They are more common in tropical and subtropical regions than in arid regions. Birds occur from sea level in dense scrub on coastal plains to 800 and 1500 m, occasionally to 2300 and 2500 m in Venezuela and Ecuador (Hilty and Brown 1986, Fjeldså and Krabbe 1990, Tostain et al. 1992, Haverschmidt and Mees 1994, Howell and Webb 1995, Stotz et al. 1996, Ridgely and Greenfield 2001). Shy and skulking most of the time, they are conspicuous when they sing on post or wire, raising the crest and spreading the alula. They sand bathe and dust bathe on the ground. Solitary in feeding, they forage in vegetation and hop on the ground, sway the body side to side, fan the large alula, perhaps to flush prey; then move in a sudden rush after prey after flashing the alula. They stand on the ground or rest on a low perch, raising and lowering the long crown, stretching and closing the wings one at a time, extending and waving the alula, and shifting the body from side to side. They also flash the alula when disturbed or frightened, and in social interactions when they raise and lower the crest (Stiles and Skutch 1989, Haverschmidt and Mees 1994, Sick 1999).

Food Insects, notably grasshoppers, caterpillars, dragonflies, cockroaches and beetles, and spiders and snails (Pelzeln 1871, Friedmann 1927, Dickey and Van Rossem 1938, Belton 1984, Haverschmidt and Mees 1994, Skutch 1999).Young, like other parasitic cuckoos, take what the foster parents bring: insects including crickets, roaches, grasshoppers and caterpillars, and some are fed fruit (Loetscher 1952).

Displays and breeding behavior Territorial birds give loud “fee-fee” songs in long series, while birds that countersing with a neighbor or that respond to the playback of either “fee-fee” or “tres pesos pide” give the “tres pesos pide” song. Birds that appear excited in approaching a playback song raise and lower the crest, hunch the body, fan the tail, and hold the wings out and down with the dark alulae extended, and sway from side to side (Smith and Smith 2000).The conspicuous behavior of erecting the long black alulae and flipping these back and forth across the paler breast gives the

186 American Striped Cuckoo Tapera naevia bird its local name of “quatro asas” or “four wings” (Skutch 1999).

Breeding and life cycle In Oaxaca, southern Mexico, in June (Binford 1989), in Panamá the birds sing from January to June (Wetmore 1968), in Colombia at S Bolívar they breed in April (Hilty and Brown 1986), in Venezuela in September (Thomas 1979). In Suriname they breed nearly all year (Hellebrekers 1942, Haverschmidt and Mees 1994). In Trinidad they sing or breed nearly all year, mainly from June to August (Belcher and Smooker 1936, ffrench 1991) and in French Guiana early in the dry season (Tostain et al. 1992). Birds in Argentina breed from October to February, mainly in November and December (Friedmann 1927, 1933, Morgensen 1927, de la Peña 1995). Brood-parasitic. Hosts of the cuckoo are small birds with covered or domed nests, especially the furnariid ovenbirds. Twenty species, mainly suboscines, are known to be parasitized: Chotoy Spinetail Schoeniophylax phryganophila, Buff-browed (Azara’s) Spinetail Synallaxis azarae, Sooty-fronted Spinetail S. frontalis, Chicli Spinetail S. spixi, Palebreasted Spinetail S. albescens, Plain-crowned Spinetail S. guajanensis, Stripe-breasted Spinetail S. cinnamomea, Rufous-breasted Spinetail S. erythrothorax, Yellow-fronted Spinetail Certhiaxis cinnamomea, Short-billed Castanero Asthenes baeri, Red-eyed Thornbird Phacellodromus erythrophthalmus, Plain Thornbird P. rufifrons, Greater Thornbird P. ruber, Buff-fronted Foliage-gleaners Philydor rufus, Whiteheaded Marsh-Tyrant Arundinicola leucocephala, flycatchers Myiozetetes spp., Plain Wren Thryothorus modestus, Rufous-and-white Wren T. rufalbus, and Black-striped Sparrow Arremonops conirostris (Friedmann 1927, 1933, Belcher and Smooker 1936, Loetscher 1952, Wetmore 1968, Kiff and Williams 1978, Salvador 1982, Sick 1993, Haverschmidt and Mees 1994, de la Peña 1995, Skutch 1999). The female cuckoo enters the bulky nest of mud

or interlocked sticks through the narrow entrance tube, and it also enters nests of Buff-fronted Foliagegleaners in burrows in the ground or a cavity in a tree (Wetmore 1968, Skutch 1999). The female enters at dawn the nest to lay, when nesting parents are foraging.When female cuckoos open the ovenbird nest near the incubation chamber, the nesting birds repair the damage. The eggs are white or bluish white to bluish green in Central America (Skutch 1999), pale blue in Colombia (Carriker, in Hilty and Brown 1998), greenish, bluish white, or greenish-blue in Suriname and Brazil (Haverschmidt 1961, Haverschmidt and Mees 1994, Dubs 1992, Sick 1993), white in Argentina (Friedmann 1927), 21 ⫻ 16 mm (Schönwetter 1964), including a record that was earlier identified as Pavonine Cuckoo Dromococcyx pavoninus which is unknown in Suriname: the eggs were white as in T. naevia (Pelzeln and Pelzeln 1910, Hellebrekers 1942). Usually one, sometimes two cuckoo eggs are laid in a host nest. Incubation period is 15–16 days (host Synallaxis incubation is 17–18 days). The host young do not survive, and although eviction has not been seen, within 24 hours of hatching, the nestling cuckoo kills the host young with its sharp bill and the foster parent removes its dead young (Morgensen 1927, Friedmann 1927, 1933, Belcher and Smooker 1936, Sick 1953a, 1981, 1993, Haverschmidt 1955, 1961,Wetmore 1968, Morton and Farabaugh 1979, Salvador 1982, Stiles and Skutch 1989, de la Peña 1993, 1995, Skutch 1999).The nestling feathers out by day 10, fledges in 16–18 days but the young bird is flightless for another week while it remains with its foster parents. The young continues to accept food from its (human) foster parent to the age of 36 days (Morton and Farabaugh 1979). Young at fledging excrete a foul-smelling liquid; behavior of nestlings was not noted. Frequency of brood parasitism: in Suriname 14 of 21 nests of Yellowthroated Spinetail had a cuckoo egg (Haverschmidt 1961).

Pheasant Cuckoo Dromococcyx phasianellus 187

Genus Dromococcyx Wied, 1832 Dromococcyx Wied, 1832, Beitrage zur Naturgeschichte von Brasilien, 4(1), p. 351. New World cuckoos with brown plumage, a crest, a slender, straight bill, broad rectrices, and upper tail coverts that extend to the end of the tail.The breast muscles and sternum are

large, and the leg muscles are weak.Type, by monotypy, Macropus phasianellus Spix 1824. The genus name refers to the terrestrial running behavior (Gr. dromos, running; kokkus, cuckoo).Two species.

Pheasant Cuckoo Dromococcyx phasianellus (Spix, 1824) Macropus phasianellus Spix, 1824, Avium Species Novae, quas in itinere per Brasilium annias 1817–20 / collegit et descripsit . . . , 1, p. 53, pl. 42. (Tonantins, Brazil) Other common names: Peixe-frito-verdadeiro (Common Fried Fish),Yasi-yatére Grande Monotypic.

back and wing coverts brown with round buff spots on tips, secondaries with buff tips, tail feathers more pointed than adult and with no subterminal black band or white tips, face with a buff eye streak that extends to nape, the throat and breast rich buff with a black border on sides, the belly and under tail coverts white and unmarked; eye-ring dusky green, iris dark gray or grayish brown, feet gray.

Description ADULT: Sexes alike, the crown chestnut with a short chestnut crest, back blackish brown, the upper back and wing coverts with white margins, alula large, wing primaries and secondaries dark brown with white tip, rump blackish, upper tail coverts nearly as long as the tail, graduated in shape, dark brown, each covert ending with a subterminal black band and a white spot, the tail long and graduated, tail feathers dark brown with a subterminal black band and white tip of the outer feathers T2 to T5; face with a white eye-streak which extends to nape, cheek blackish brown; the chin whitish, the breast buffy white with fine blackish spots and streaks formed by dark rachis and dark terminal barbs, belly and under tail coverts white and unmarked, under wing coverts white, base of primaries white and a white bar across center of middle primaries; bare skin around eye yellow-green to blue-green, bare lores bluish green to dull greenish gray, iris brown to yellow, bill blackish above, below gray, feet grayish brown. JUVENILE: Crown black with fine buff spots in a short black crest, upperparts sooty brown, the upper

NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, CM, DMNH, FMNH, LSU, MCZ, MVZ, NMW, ROM, SMF, SMT, UMMZ, USNM, ZSM.

Subspecies None. Regional texts on birds in Mexico and Central America often recognize a subspecies D. p. rufigularis Lawrence, 1867. The form was described from a bird in juvenile plumage, and the type (AMNH 44460) was compared only with birds in adult plumage. Neither juveniles nor adults differ in size or in plumage between geographic regions.

Measurements and weights Wing, M (n ⫽ 22) 154–176 (166.2 ⫾ 5.9), F (n ⫽ 18) 153–173 (162.8 ⫾ 5.6); tail, M 173–242 (209.4 ⫾ 17.0), F 177–248 (203.1 ⫾ 19.0); bill, M 17.7–25.2 (20.4 ⫾ 2.0), F 17.5–23.6 (20.0 ⫾ 1.6); tarsus, M 27.0–35.7 (31.5 ⫾ 2.8), F 27.3–36.9 (30.9 ⫾ 3.2) (AMNH, ANSP, UMMZ). Weight, M (n ⫽ 4) 78–98 (85.0), F, laying (n ⫽ 1) 98.1 (AMNH, ANSP, UMMZ).

188 Pheasant Cuckoo Dromococcyx phasianellus Wing formula, P7 ⬎ 8 ⫽ 6 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⫽ 10.

Field characters Overall length 36 cm. Long-tailed cuckoo of the forest floor, with a small head and thin neck, a long fan-shaped tail with white spots, a chestnut crest and a white line through the eye, white with dark streaks below. Juvenile has a blackish crown with fine spots, buff scallops on upperparts, and rich buff on the throat and breast. The tail feathers are broader than in the Pavonine Cuckoo (D. phasianellus, adult and juvenile width 28–34; D. pavoninus, width 20–24), and the tail tip is less marked with white.

Voice Clear whistles, “whee whee wheerr-rr”, the first notes a series of 2–4 successively higher pitched whistles rising from 1.8 to 2.0 kHz, the last note a tremulous trill, the song lasting about 1.5 sec. Other calls include a series of rattling, clucking or growling “grrr” notes (Dickey and Van Rossem 1938, Wetmore 1968,Willis and Eisenmann 1979, Belton 1984, Hilty and Brown 1986, Stiles and Skutch 1989, Hardy et al. 1990, Howell and Webb 1995, de la Peña and Rumboll 1998). The bird approaches when it hears a good imitation of its whistle calls.

Range and status Southern Mexico (Guerrero, Puebla, Oaxaca, Veracruz, Chiapas and the Yucatan Peninsula) through the Caribbean and Pacific slopes of Central America (mainly on Caribbean slope, in Belize, Guatemala, northern Honduras and Nicaragua, and eastern El Salvador and adjacent southeastern Honduras) and Panamá, and in South America east of the Andes from Colombia (middle Magdalena Valley, eastern base of Andes, in northeast in Arauca and Vichada (Río Meta) and in the extreme southeast in Amazona (Leticia), Venezuela (Zulia, Carabobo, Aragua, Guárico, Sucre, NE Bolívar), Brazil (Roraima, Para (Aramanaí), Amazonas (Rio Madeira and Borba), Maranhão (Tabocas), Piauí, Bahia (Barra, Bomfin), Goiás (Araguay), Minas Gerais, Mato Grosso (including Engenho do

Gama, Descalvado), Mato Grosso do Sul, Paraná, Rio de Janeiro, São Paulo and Rio Grande do Sul), Amazonian Ecuador (E Sucumbíos, E Napo, E Pastaza, E Morona Santiago) and eastern Peru (Ucayali, Rio Apurimac in Cuzco), to N Bolivia (Rio Beni, Rio Chapare), Paraguay (scarce in Chaco, present in Paraguarí, Ñeembucú and Cordillera) and NE Argentina (Misiones) (Pelzeln 1871, Ridgway 1916, Hellmayr 1929, Pinto 1938, Meyer de Schauensee and Phelps 1978, Hilty and Brown 1986, Contreras 1993, Hayes 1995, Howell and Webb 1995, Parker and Goerck 1997, AOU 1998, Cappeer et al. 2001, Ridgely and Greenfield 2001). In Ecuador it is seen near Kapawi Lodge along the Río Pastaza near the Peruvian border. Unknown in the Guianas. Resident. Widespread, uncommon to rare and local, secretive, solitary.The species was not seen after 1971 on Barro Colorado I, Panamá, where it was one of the first birds to disappear when the island lost its cuckoos and the possibility of immigrants due to deforestation on the nearby mainland (Willis and Eisenmann 1979).

Habitat and general habits Humid understory of young woodland, tropical lowland evergreen forest, flooded tropical evergreen forest, tropical deciduous forest, thickets and undergrowth, low-altitude cloud forest, forest borders and secondary woodland, lake margins with grassy edges, thick growth of vines, Cecropia and dying trees; mainly lowlands, occurring from sea level to 700 m in Ecuador), 1200 m in Panamá and 1600 m in Colombia (Wetmore 1968, Terborgh et al. 1984,

Pavonine Cuckoo Dromococcyx pavoninus 189 Hilty and Brown 1986, Sick 1993, Stotz et al. 1996, 1997, de la Peña and Rumboll 1998, Ridgely and Greenfield 2001). Solitary, secretive and skulking, they occur on the ground or near the ground, less often in the crowns of tall trees. Although the cuckoos are terrestrial, walking and wagging the tail, the tail and sternum are large and well muscled and the leg muscles are weak (Dickey and Van Rossem 1938, Wetmore 1968). The foraging bird bobs the body and makes rattling noises, vibrating the wing feathers and bill while it fans the tail downward and brushes the litter, then the bird pauses, rushes forward flicking its wings, and pecks in the litter (Sieving 1990).

Food Insects, including large grasshoppers and cicadas, beetles; small vertebrates including lizards (anoles, gekkos) and nestling birds (Pelzeln 1871, Wetmore 1968, Stiles and Skutch 1989).

Displays and breeding behavior Territorial, a singing adult raises head and crest, speckled breast feathers puffed out, alula and outer

primaries partly extended to show white spots, with upper tail coverts strongly arched. Two birds walk in parallel in display within 1 m of each other (Sieving 1990).

Breeding In Oaxaca, Mexico, from April to June (Binford 1989), in Bahia and Mato Grosso, Brazil, in November and December (Naumburg 1930, Wetmore 1968), and calling daily at dawn in Matogrosense, Paraguay, in September (Capper et al. 2001). Brood-parasitic. Hosts are small birds, mainly tyrannid flycatchers, that lay in either open or closed nests. In Mexico they use Yellow-olive Flycatcher Tolmomyias sulphurescens (Wilson 1992), in Brazil and Argentina the Eye-ringed Flatbill Rhynchocyclus brevirostris, Myiozetetes sp. and Pied Water-Tyrant Fluvicola pica and Barred Antshrike Thamnophilus doliatus (Sick 1993). Eggs are dull white or pale buff, with a wreath of rufous spots on the large end (Naumburg 1930, Sick 1953a,b, 1993, Friedmann 1964b, de la Peña 1986), 25 ⫻ 14.5 mm (oviduct egg, Naumburg 1930). Incubation and nestling periods are unknown.

Pavonine Cuckoo Dromococcyx pavoninus Pelzeln, 1870 Dromococcyx pavoninus Pelzeln, 1870, Ornithologie Brasiliens, 3, p. 270. (Araguay, Engenho do Gama and Arimani, Brazil). Other common names: Peacock Cuckoo, Yasiyatére Chico Monotypic.

Description ADULT: Sexes alike, crown and face rufous, short rufous crest, back dark brown with feathers edged in white, alula large, wing coverts edged whitish, wing dark gray brown, rump blackish, elongated and narrowly graduated black upper tail coverts with a small white spot at the tip, coverts cascading nearly the length of the tail, tail long and graduated, brown with broad white tip, the tail from below

gray with a subterminal black band and white tip, face with ear patch dark rufous and a deep buff eye streak extending to the nape; the throat and breast unspotted rufous, belly and under tail coverts white, under wing coverts white; orbital skin greenish gray with an inner ring yellow and black, iris brown, bill black above and gray below, feet gray to gray brown. JUVENILE: Crown dark brown, back dark brown without white edges of feathers, wing coverts and wing much as in the adult, upper tail coverts and tail long without terminal markings of white, cheek gray brown, eye streak light buff, chin white, throat gray, breast brown gray, belly whitish. NESTLING: Undescribed.

190 Pavonine Cuckoo Dromococcyx pavoninus SOURCES: AMNH, ANSP, BMNH, CM, DMNH, LSU, NMW, SMF, UMMZ, USNM, ZSM. A subspecies D. p. perijanus Aveledo H. and Ginés 1950 from the Upper Río Negro, Zulia,Venezuela, was described as dark on the crown. The holotype (Phelps collection no. 1172, in AMNH) is a juvenile with a few spotted upper tail feathers, the crown like the dark crown of juveniles from other areas.

Measurements and weights Wing, M (n ⫽ 13) 129–145 (134.4 ⫾ 4.7), F (n ⫽ 5) 126–142 (133.8 ⫾ 3.4); tail, M 132–174 (150.4 ⫾ 13.0), F 135–170 (152.8); bill, M 18.3–22.2 (19.6 ⫾ 1.4), F 18.4–19.7 (19.1); tarsus, M 26.6–33.2 (28.2 ⫾ 1.2), F 26.2–26.9 (26.5) (AMNH, ANSP, BMNH, CM, DMNH, LSU, USNM). Weight, M (n ⫽ 1) 50, F (n ⫽ ?) 45.2; U (n ⫽ 2) 48–49 (AMNH, Sick 1993, Magalhães 1999). Wing formula, P7 ⫽ 6 ⬎ 5 ⬎ 8 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 28 cm. Long-tailed cuckoo of the forest floor, small head and thin neck, barred blackish back, long graduated tail with white spots, rufous crest and buff line through the eye, throat and breast unspotted rufous. Juvenile head gray-brown, eye streak white and back unmarked. Tail feathers narrower than in Pheasant Cuckoo (D. pavoninus, feather width 20–24; D. phasianellus, feather width 28–33).

Voice Song a high whistle “fee fee, feefee,” the first note lowest, second and fourth notes a half tone higher. Song is higher pitched (2.4–2.6 kHz), lasts about 2 sec, and lacks terminal tremolo that occurs in song of Pheasant Cuckoo (Hardy et al. 1990).They call before dawn (Boesman 1998).

Range and status South America east of the Andes from Colombia (Tierra Nueva, Sierra Negra SE of Fonseca 4000’: USNM 368717), Venezuela (subtropical western slope of the eastern Andes in NW Zulia, Aragua, S Bolívar, S Amazonas), Guyana (mainly near coastal

rivers: Abary River, Kartabo, Bartica, Georgetown, Annai River, Iwokrama Reverve on Essequibo River, Moraballi Creek, Ituribisi River, Supenaam River, Bonasika River; inland at Mt Roraima), French Guinea (Saut Tamanoir on Fleuve Mana), E Ecuador (Río Suno in W Napo, observed on the slopes of Cordillera del Cóndor), E Peru (Ucayali), Bolivia (La Paz, Santa Cruz, Cochabamba), Brazil (Amazonas (Manaus, Rio Branco, Rio Madeira), Maranhão, Pará (Obidos, Upper Rocaua), Goiás, Minas Gerais, Mato Grosso (Chapada), São Paulo (Botucatú, Itapura, Iguape, Albuquerque Lins), Paraná and Rio de Janeiro, Paraguay (mainly in the east) and NE Argentina (Misiones). Resident. Solitary, uncommon, their distribution is discontinuous; there is one record in Colombia and none in Suriname (Pelzeln 1871, Chubb 1916, Pinto 1938, Snyder 1966, Short 1975, Meyer de Schauensee and Phelps 1978, Remsen and Ridgely 1980, Dubs 1982, Remsen and Traylor 1983, Hilty and Brown 1986, Sick 1993, Haverschmidt and Mees 1994, Hayes 1995, Stotz et al. 1996, Cohn-Haft et al. 1997, Boesman 1998, de la Peña and Rumboll 1998, Kirwan and Sharpe 1999, Capper et al. 2001, Ridgely and Greenfield 2001, Salaman et al. 2001).

Habitat and general habits Understory of humid lowland tropical evergreen forest, montane evergreen forest, tangled thickets with dense secondary woodland, bamboo, heavy brush, transitional forest, seasonally inundated forest, occur with abundant Heliconia and Ficus, usually near water. Live in lowlands to 1600 m, occasionally to

Lesser Ground-cuckoo Morococcyx erythropygus 191 1900 m in Venezuela, in Ecuador c. 400 m (Terborgh et al. 1984, Hilty and Brown 1986, Haverschmidt and Mees 1994, de la Peña and Rumboll 1998, Ridgely and Greenfield 2001). Solitary. Forage on the ground and in forest understory.

Food Insects, mainly orthoptera, caught on the ground (Chubb 1916, LSU).

Breeding In Guyana a female cuckoo had a large ovum (7 mm) in July (ANSP 187796). Brood-parasitic. Hosts are several species of suboscine passerines with closed or bag-shaped nests (Tyrannidae:

Ochre-faced Tody-Flycatcher Todirostrum plumbeiceps, Eared Pygmy-tyrant Myiornis auricularis, todytyrants Hemitriccus spp; Thamnophilidae: Plain Antvireo Dysithamnus mentalis). It is unknown how the cuckoo gets its eggs into a small covered nest it cannot enter, such as the nest of a 6 g todyflycatcher: perhaps the cuckoo lays while it holds onto the nest outside the nest entrance (Sick 1993). Eggs are rose white with purplish spots, 21 ⫻ 15 mm (Giai 1949, Neunteufel 1951, Sick 1953a,b, 1993, Schönwetter 1964, de la Peña 1986, Haverschmidt and Mees 1994). Incubation and nestling periods are unknown. The host young disappear after the cuckoo hatches, a circumstance pointing to infanticide by the nestling cuckoo.

Genus Morococcyx Sclater, 1862 Morococcyx Sclater, 1862, Catalogue of Collection of American Birds, p. 322. Cuckoo with a short, decurved bill and long legs. Type, by monotypy, Coccyzus

erythropyga Lesson 1842.The name refers to the harlequin pattern of the head plumage (Gr. moros, clown, silly, foolish; kokkux, cuckoo). One species.

Lesser Ground-cuckoo Morococcyx erythropygus (Lesson, 1842) Coccyzus erythropyga Lesson, 1842, Revue et Magasin de Zoologie Pure et Appliquee (Paris) 5, p. 210. (San-Carlos, Centre Amérique ⫽ La Unión, El Salvador) Other common names: Rufous-rumped Cuckoo Polytypic. Two subspecies. Morococcyx erythropygus erythropygus (Lesson 1842); Morococcyx erythropygus mexicanus Ridgway 1915.

Description ADULT: Sexes alike, crown brown streaked with black and tipped buff, back grayish brown, wing brown with bronzy gloss, lower back and rump black with rufous tips, loose-webbed upper tail coverts rufous, long brown tail,T1 unmarked,T2 to T5 with subterminal black bar extending onto lateral vane and with buff tip, face with a buffy white superciliary streak which broadens to pale gray or

buffy gray over the ear; underparts, throat and breast rufous buff, lower belly and under tail coverts dark cinnamon; bare skin above and behind the eye blue (darker below the eye) bordered with black feathers, eye-ring and bare skin in front of eye yellow, iris brown, bill decurved, black above and yellow below, long legs yellow-brown. JUVENILE: Duller above, feathers scaled with pale buff edges and tips from crown to rump, rectrices more pointed (less rounded) than in adult, T1 unmarked brown, T2 to T5 brown without a subterminal black bar and with a small buff tip; underparts rufous smudged dusky; bare skin around eye gray, iris brown, bill above brown below lighter. NESTLING: Undescribed.

192 Lesser Ground-cuckoo Morococcyx erythropygus SOURCES: AMNH, ANSP, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM.

Subspecies and geographic variation Morococcyx erythropygus mexicanus Ridgway, 1915; larger and paler, crown uniformly colored or indistinctly streaked, black tail marks indistinct; Pacific slope of tropical western Mexico from S Sinaloa south to Guerrero and Isthmus of Tehuantepec in Oaxaca and to Arriaga in SW Chiapas; Morococcyx erythropygus erythropygus (Lesson, 1842); smaller and darker, crown darker than the back and streaked with dark central shafts, black tail marks conspicuous when seen from below, tail tips buff or white; S Mexico in Chiapas east of the Isthmus of Tehuantepec, and Pacific Central America south to NW Costa Rica, and the arid interior valleys on Caribbean slope of Guatemala and Honduras. Plumage varies in color of the rufous underparts, the darker birds in El Salvador, paler eastward to Costa Rica; the gradient between rufous birds on the east and paler birds on the west side of the Isthmus of Tehuantepec is abrupt.

Measurements and weights M. e. mexicanus: wing, M (n ⫽ 9) 98–107 (101.3 ⫾ 2.7), F (n = 8) 96–105 (100.6 ⫾ 3.0); tail, M 123– 148 (135.9 ⫾ 8.2), F 125–143 (136.1 ⫾ 6.7); bill, M 21–24 (23.9 ⫾ 2.9), F 22–26 (23.7 ⫾ 1.6); tarsus, M 30–38 (34.6 ⫾ 2.8), F 30.5–39 (34.8 ⫾ 3.4) (AMNH, FMNH, UMMZ; M. e. erythropus: Guatemala to Costa Rica, wing, M (n ⫽ 22) 93–105 (98.7 ⫾ 3.6), F (n ⫽ 15) 90– 102 (96.9 ⫾ 3.8); tail, M 126–140 (130.2 ⫾ 6.7), F 120–142 (128.7 ⫾ 7.4); bill, M 19.4–25.2 (22.2 ⫾ 1.6), F 20–23.5 (21.8 ⫾ 1.0); tarsus, M 30.2–36.4 (33.6 ⫾ 1.8), F 31.4–36.0 (33.5 ⫾ 1.6) (FMNH, UMMZ). Weight, M. e. mexicanus: M (n ⫽ 5) 58–70.5 (63.1), F (n ⫽ 2) 62.9–86.8 (68.0); M. e. erythropus: M (n ⫽ 10) 53.1–66.2 (61.1), F (n ⫽ 7) 56–76 (65.3) (AMNH, MVZ, LSU, UMMZ). Wing formula, P7 ⫽ 6 ⬎ 8 ⫽ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⫹ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 25 cm. Greyish brown cuckoo with rufous buff underparts, blue skin around the eye bordered with black, and a pale mark over the ear.

Centropus rectunguis Wing, M (n ⫽ 4) 156–170 (165.5), F (n ⫽ 2) 166– 186 (176); tail, M 192–204 (198), F 194–238 (216); bill, M 35–37 (36), F 34–42 (37); tarsus, M 42–46 (44.3), F 45–52 (48.5); hallux claw, M 8–12 (10.0), F 9–19 (14) (AMNH, BMNH, ZRC).

Voice A loud mellow curlew-like whistle “teeeee,” flat at first then rising in the second half of the call and ending with a drop, the whistle mainly around 1.4 kHz and rising to 2 kHz, and lasting 0.6 sec. Variations include a clear trilled whistled series with notes lower and more widely spaced, the series introduced by 2–3 clear ascending whistles. Second call type is a soft burry whistle “whirrr” of four rising notes, the second two notes louder and higher-pitched than the first and last notes, the phrase lasting 0.3 sec. Female clacks the mandibles when disturbed (Stiles and Skutch 1989, Hardy et al. 1990, Howell and Webb 1995).

Range and status Southern Mexico and Central America. On the Pacific slope of western Mexico they occur from southern Sinaloa south to interior Mexico in the Balsas River region in Michoacan and Guerrero,

Greater Roadrunner Geococcyx californianus 193 and along the coast in Oaxaca and Chiapas, and in western Central America from Guatemala, El Salvador and Nicaragua to northwestern Costa Rica (Río Grande de Tárcoles).They also occur in the arid interior valleys on the Caribbean slope of Guatemala (Motagua) and Honduras (Quimistán, Sula, Comayagua, Aguán) where the population range is continuous with birds of the Pacific slope (Ridgway 1916, Hellmayr 1929, Howell and Webb 1995, AOU 1998). Resident. Fairly common, and locally common in Guanacaste, Costa Rica.

furtive and skulking, yet inquisitive and not shy (Stiles and Skutch 1989). Usually alone or in pairs, they behave like miniature roadrunners in terrestrial behavior, but walk slowly and deliberately, and freeze in position when alarmed (Howell and Webb 1995).

Food Insects, mainly grasshoppers (Dickey and Van Rossem 1938).

Breeding and life cycle Habitat and general habits Arid lowland scrub, tropical deciduous forest edge in semi-arid scrub and woodland, thorny thickets on arid slopes, agave plantations; sea level to 1500 m, locally to 1800 m in Guatemala (Stotz et al. 1996). Terrestrial, the birds forage in understory and near the edge of cultivation, and walk or hop on bare ground and among leaves on the ground with head and leg movements like a chicken, pecking food from the ground (Rowley 1984). They leap into foliage and walk along branches much like a dove (Berger 1960).They are

In Oaxaca, Mexico, they breed in the rainy season, from May to November (Rowley 1984, Binford 1989, UMMZ), in Costa Rica from February to May (Stiles and Skutch 1989). Nest is a shallow bowl of sticks and leaf petioles and rachises, lined with dead leaves, built on the ground. Eggs are chalky or smooth white, 27 ⫻ 21 mm, clutch 2 (1–3) (Skutch 1966, Rowley 1984, Howell and Webb 1995). Both sexes incubate, each taking its turn for 2–4 hours at a time before the mate takes over (Skutch 1966, 1983). Incubation and nestling periods are unknown.

Genus Geococcyx Wagler, 1831 Geococcyx Wagler, 1831, Isis von Oken, col. 524. Large, slender terrestrial cuckoos with streaked plumage, slender build, a crest, long legs and a long tail, living in arid and semi-arid regions of North and Central

America. Type, by monotypy, Geococcyx variegata Wagler ⫽ Saurothera californiana Lesson (Peters 1940). The name refers to the terrestrial habits of the bird (Gr. geo, earth; kokkux, the cuckoo).Two species.

Greater Roadrunner Geococcyx californianus (Lesson, 1829) Saurothera Californiana (Lesson, 1829), Oeuvres Complètes de Buffon, 6, p. 420. (Californie ⫽ San Diego, California). Monotypic.

Description ADULT: Sexes alike, above the crown and nape black with whitish spots and streaks, shaggy blueblack crest, streaked brown with bronze gloss,

upper back blackish brown with buff edges of the feathers forming whitish streaks, wing coverts blackish brown with buff and white, wing dark brown with two white stripes across the primaries, tip of outer vane of primaries edged white, lower back and rump brownish gray, rectrices broad and truncate, tail black with white outer edge on rectrices T1 to T5 and with broad white oval tips on T3 to T5; underparts whitish with brown streaks

194 Greater Roadrunner Geococcyx californianus and a black shaft, chin whitish, flanks whitish to whitish gray, under tail coverts light gray; bare skin behind eye blue and orange (behind eye males are usually white, females blue), the colorful skin normally concealed by feathers is exposed in display, eye-ring blue, iris brown with yellow ring around the pupil, bill blackish, tarsi brown to gray. JUVENILE: Plumage without gloss, feather tips notably the outer primary coverts and the rectrices more narrow and tapered than in adults, irregular edge of white patch on rectrices T3 to T5; postorbital skin as in adult female, white patches on tips of outer primary coverts have inverted U shape (less concave in adult); inner iris with a gray ring around pupil, outer iris brown. NESTLING: Nearly naked, skin black (throat whitish) with short (1–7 mm) whitish hair-like down; gape flange pink, mouth red with upper palate white, a raised area of white papillae on each side of palate and on the rear edge of the tongue, and tongue with a black tip, iris dark grayish brown without yellow ring; legs and feet are black at hatching and gray by 11 days. SOURCES: AMNH, CM, CMNH, FMNH, LSU, MCZ, MVZ, ROM, UMMZ, USNM.

Measurements and weights Arizona, New Mexico, Oklahoma and Texas south to Mexico (Sonora and Tamaulipas to Michoacan): Wing, M (n ⫽ 16) 163–190 (178.3 ⫾ 8.3), F (n ⫽ 12) 158–186 (170.7 ⫾ 9.0); tail, M 256–302 (283.6 ⫾ 14.3), F 260–308 (275.5 ⫾ 14.4); bill, M 43–53 (48.1 ⫾ 3.2), F 41–51 (45.5 ⫾ 3.2); tarsus, M 56–66 (61.4 ⫾ 2.7), F 56–64 (59.5 ⫾ 2.8) (UMMZ); California, Arizona and New Mexico: wing, M (n ⫽ 21) (185.5 ⫾ 18.4), F (n ⫽ 11) (171.6 ⫾ 10.0); tail, M (301.6 ⫾ 18.3), F (282.1 ⫾ 17.0); bill, M (36.7 ⫾ 2.3), F (35.4 ⫾ 1.2); tarsus, M (61.3 ⫾ 1.8), F (59.8 ⫾ 1.6) (Hughes 1996b). Weight, California: M (n ⫽ 8) 275–430 (344.0), F (n ⫽ 6) 264–352 (308.5) (MVZ); Texas, M (n ⫽ 6) 187–333 (270.0) (UMMZ); Oklahoma: early winter (n ⫽ 5) 412–588 (465), midwinter (n ⫽ 3) 300–325 (314.7) (Geluso 1969, 1970a, Sutton 1973).

Wing formula, P7 ⫽ 6 ⫽ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10. Oberholser (1974) noted that birds from east and central Texas south through eastern Mexico to Veracruz are smaller on average, and described a subspecies G. c. dromicus from Brownsville, Texas. These birds overlap in size with birds elsewhere in the species range (Browning 1978, Hughes 1996b).

History The first recorded description of the roadrunner was by a Franciscan priest in California in 1790, as reported by Coues (1900). “The Churcha is a kind of pheasant which has a long bill, dark plumage, a handsome tail and four feet. It has these latter facing outward in such fashion that when it runs it leaves the track of two feet going forward and two going backward.” Apparently “feet” replaced “toes” in a translation of this account.

Field characters Overall length 56 cm. Large terrestrial cuckoo with long legs and long tail, cocked up or hanging or held straight behind as bird runs over the ground. Greater Roadrunner differs from Lesser Roadrunner G. velox in having a breast more uniformly streaked with brown (not boldly streaked black), a lack of buff on the sides, a gray (not chestnut) back and rump, and pale under tail coverts.The colors of the bare skin behind the eye are exposed when the crest is raised. (Plate 2), at other times the full colors are not seen (Figure 4.10)

Voice Mournful dove-like “coo” notes given in long series “co-coo-coo-coo-cooooooo” at low frequency, slow and resonant, the notes dropping in pitch, the first notes about 0.6 kHz and the last notes 0.5 kHz, in a series the first note short and the later notes longer, the song lasting 3 or 4 sec, heard to about 250 m and given by the male; sometimes gives a short coo or paired coo notes in a soft two-note series “cooocooo.” Other calls are growl-coos, “whir” calls, whines, and bill clacks and rattles “trrrt”. Another sound is produced by the wings pulled inward towards the body in a series of four to eight loud

Greater Roadrunner Geococcyx californianus 195 counties of California with increasing agricultural and urban development and human disturbance (Ridgway 1916, Grinnell and Miller 1944, Emlen 1974, Garrett and Dunn 1981, Tweit and Tweit 1986, McCaskie et al. 1988, Hughes 1996b). Roadrunner population numbers showed no significant change over most of the range of the species in the period 1966–1993 (Price et al. 1995).

Habitat and general habits

“pop” sounds. Females and males often engage in mutual calling.The female gives a bark, and the male responds with a “growl-coo”, often while she is on the nest during egg-laying and incubation.The male also gives a “growl-coo” in response to his female’s “bark” while they are foraging apart from each other (Whitson 1971, 1975, 1983, Hardy et al. 1990, Hughes 1996b, K. Groschupf ).

Range and status North America, in southwestern United States from the Central Valley of California, extreme S Nevada, southern Utah, Kansas and Oklahoma south to the Gulf coast of Texas and east to W and NE Louisiana, Arkansas and extreme S Missouri, and in Mexico in Baja California, Sonora, Sinaloa, Durango, Zacatecas, Jalisco, eastern Michoacán, Puebla, Veracruz and northern Tamaulipas (Ridgway 1916, Grinnell and Miller 1944, Oberholser 1974, Howell and Webb 1995, Hughes 1996b, AOU 1998). Resident. Common to fairly common, roadrunners are most numerous in SE California, southern Arizona (the Chihuahuan desert) and Texas west of Pecos River and south of Edwards Plateau (Price et al. 1995). Their range expanded north and eastwards through the twentieth century in California, Kansas and Oklahoma and across the Ozark Mts into the Mississippi River basin (James and Neal 1986, Hughes 1996b). Land clearance, overgrazing and the invasion of shrubs may have caused eastern range expansions. Their range has decreased in regions where their native habitat was developed for agricultural and residential use. Roadrunners have disappeared in the Central Valley and southern

Semi-arid and arid lowland scrub, desert scrub, arid montane scrub, widely dispersed in dry open country with scattered brush, mesquite, palo verde, creosote, cholla cactus, prickly-pear cactus, upper oak and piñon pine scrub, in yucca and shortgrass plains of north Texas panhandle, also in tall pines and magnolias in east Texas, clearings in farms and dry scrubby woods ( James and Neal 1986, Stotz et al. 1996). They occur mainly in arid and semi-arid regions that have at least 140 clear sunny days. The northern extent of the range is limited by prolonged snow cover (Beal 1981). In Mexico, roadrunners occur from the lowlands to as high as 2900 m in the Cerro San Andres, Michoacán (UMMZ). Greater Roadrunners, cuckoos of arid scrub and desert, are the best known of any cuckoos for their morphological and physiological adaptations to their habitat. They have dorsal skin with large black patches of melanin, a high body weight, a salt-secreting nasal gland and an ability to reabsorb water through the cloaca (Calder and Bentley 1967, Ohmart et al. 1970a,b, Ohmart 1972, Dunson et al. 1976). They also have an ability to cool the body by evaporating water through the respiratory system and the skin (Lasiewski et al. 1971), to lower the body temperature at night, and to use solar radiation by basking in the early morning, exposing the unfeathered and pigmented black dorsal skin which absorbs the sun’s radiant heat (Ohmart and Lasiewski 1971, Whitson 1983). In cool times the birds sunbathe in early morning by spreading the back feathers, drooping the wings and exposing the black dorsal skin to the sun. Sunning behavior reduces the energy cost to raise its body temperature from the night-time level; the savings can be as much as 60% (Ohmart 1989). Roadrunners have fat deposits under their skin in winter, and some

196 Greater Roadrunner Geococcyx californianus males also have fat deposits under the skin when they breed (Vehrencamp 1982a). Metabolic rate is about the same as that of other birds of the same body size. At air temperatures above 36°C their metabolic rate increases and at 44.3°C it is 31% above the standard metabolic rate. Roadrunners maintain their basal metabolic rate at air temperatures as low as 9°C when they bask in the sun, but without solar radiation their metabolism begins to rise if air temperature falls below 27°C, and metabolic rate increases with lower temperature below this level (Calder and SchmidtNielsen 1967). During daytime conditions of extreme heat the adult roadrunners crouch, spread the wings and lift the upper body feathers, allowing air to flow between the layers without exposing the down feathers to the sun (Meinzer 1993).They also remove heat when they flutter their gular region, pant, and extend the wings to expose lightly feathered areas under the wing (Calder 1968a), and their heat exchange in arteries and veins cools the brain (Kilgore et al. 1976). By controlling their behavior the roadrunners reduce their food demands. Nestling roadrunners warm themselves by moving into a patch of sunlight in the nest; the skin (except the chin and vent) is black and absorbs solar radiation (Ohmart and Lasiewski 1971), and they cool themselves by fluttering their gular skin and aligning their bodies to the shade. Water lost in gular flutter and cutaneous evaporation is balanced by salt-secretion of the nasal glands and by the parents regurgitation of a clear liquid (water?) to the nestlings. Adults conserve water by reabsorption through the lining of the rectum, ceca and cloaca. Roadrunners excrete salt through the nasal glands in front of the eye; the glands are more prominent in nestlings than in adults. Roadrunners can maintain their body weight with water from their diet of reptiles and rodents, but they drink when water is available. Breeding adults eat their nestlings’ urine and feces, which are enclosed in a fecal sac.The nestlings’ excreta contains more water than adults’ excreta, and adults meet part of their daily water requirements through parental care (Ohmart 1973). In addition, adult roadrunners rest inactive in the shade during

the hot mid-day, and hunt in the early morning and late afternoon when the sun is low and their insect and lizard food is active. Resident, occur alone or in pairs. Behavior has been watched by following wild or tame handreared roadrunners on their daily rounds in the field (Sutton 1915, 1922, Marshall 1957, Whitson 1971, 1983). Roadrunners run in the open or under cover on bare ground, as they cover a large foraging area quickly and secretively. They run faster than a human, as fast as 30 km / h (Kavanaugh and Ramos 1970, Ohmart 1989).They feed in the cool hours of the morning and late afternoon, and rest in shade in mid-day heat (Calder 1968a). Occasionally they fly to a tall perch to sing or to a nest, and they ascend trees almost vertically by gaining momentum from running down another branch. In a tree they climb and scramble where the vegetation allows, moving by leaps and short flights, run along the horizontal branches, and glide down slopes just over the scrub. At night they roost high in vegetation. Terrestrial, they run after prey, leap at flying insects, ambush sparrows, wait at nectar feeders and take insects and hummingbirds. Roadrunners toss and batter lizards and snakes against a stone, and they attack scorpions at their tail (Sutton 1913, Calder 1968c, Mayhew 1971, Beal and Gillam 1979).A pair sometimes cooperate in attacking a snake (Whitson 1983, Meinzer 1993). Territory size averages 0.5 km2. Population density on rocky slopes in southern California is 0.65 birds / km2 and is lower in other desert habitats ( Weathers 1983); in coastal south Texas it is estimated at 2.5–3.1 birds / km 2 (Folse and Arnold 1976, 1978). Adult annual survival is at least 60% (Folse 1974). Roadrunners breed in their first year (Smith 1981).There are few banding recoveries.

Food Opportunistic carnivore: insects (grasshoppers, crickets, cicadas, caterpillars, beetles), scorpions, centipedes, spiders, tarantulas, toads, lizards, small snakes including rattlesnakes, small birds and eggs, mice, roadside carrion, young ground squirrels,

Greater Roadrunner Geococcyx californianus 197 young rabbits and young bats fallen from cave ceiling or colliding with the cave wall (Bryant 1916, Gorsuch 1932, Sutton, in Bent 1940, Herreid 1960, Zimmerman 1970, Shetlar 1971, Oberholser 1974, Bleich 1975, Binford 1979, Laycock 1985, Hughes 1996b). Uses bill to turn over dried mud crust on fields, and feed on the exposed crickets ( Jaeger 1947). Fruits and seeds are taken as well (prickly pear cactus Opuntia, sumac Rhus integrifolia) (Bendire 1895, Hughes 1996b). In winter, feeds on insects and grain (Geluso 1970b, Sutton 1973).

Displays and breeding behavior A pair remains on its territory all year and from year to year—one pair remained in the same site over five years. Sometimes a pair move together to a new territory and renest, while other birds move independently of the mate (Folse 1974). The pair spends time foraging together, calling back and forth, before the birds show sexual interest in each other. In courtship, one bird runs after the other on the ground, often for hours; both birds stop and rest between these chases.The chaser runs and lunges at the other bird with its wings and tail raised and fanned, while both birds give “clack” calls.The male gives “coo” calls from an elevated perch. Both sexes pick up sticks and pass them to the mate or drop them at the mate’s feet. Before they copulate the roadrunners display with a “prance display.” The male approaches the female in short, quick burst of speed, holds food in his bill, and wags his tail back and forth. He runs towards and away from the mate with his wings and tail lifted, then lowers the wings and brings them inward with a “pop.” In display he holds his tail over his body and then gradually lowers the tail, exposes the postorbital bare skin, erects the crest, and sleeks the contour feathers, and lifts the wings; the behavior lasts more than two minutes. In a “tail-wag display” the male wags the tail from side to side, while he bows then slowly lifts his head, faces the female and gives a “whirr” call. He usually holds food or plant material in his bill. After this he jumps into the air and leaps over the female or mounts her from behind. In courtship feeding, the male gives mice, small birds, snakes and lizards to his mate. He holds the food in his bill and presents it to the

female as they copulate. Then he dismounts, the pair walk away from each other, flicking the tail, and the female eats the offered food, or she feeds it to her young if copulation occurs after the young have hatched (Rand 1941b, Calder 1967a,Whitson 1971, 1975, Hughes 1996b). Roadrunners have bred successfully in captivity (Muller 1971,Whitson 1971, 1975, Smith 1981).

Breeding and life cycle In southern California in low elevations of the Lower Sonoran Life Zone the birds begin nesting in late February, and in the upland deserts of the Upper Sonoran Life Zone they nest a month later (Miller and Stebbins 1964). In southern Arizona, a few nest in late March, more nest from mid-April to mid-June, and birds nest again from late July to midSeptember with a pause in the hot dry summer. Nesting after the summer rains varies with the rains of the year (Ohmart 1973, 1989). In Texas they breed from March to October (Van Tyne and Sutton 1937, Whitson 1971, Oberholser 1974, Hughes 1996b), in Oklahoma from April to July or August (Sutton 1967, Baumgartner and Baumgartner 1992), in Sonora from May to July (Russell and Monson 1998). The nest is built 1–3 m above ground in a bush, low tree, thicket, or cactus clump, an open bulky flat platform of sticks, about 30 cm wide, often placed in shade, lined with leaves, snakeskins, cattle dung and mesquite seed pod debris. Birds continue to add material to the nest during incubation and after hatching. Occasionally they use an old nest or even an active nest of other species. Eggs are white, the inner layer of eggshell dull white, outer layer glossy sometimes with a yellowish film, 39 ⫻ 30 mm (Bent 1940). The eggs are laid at intervals of 2 days, the timing is variable. Clutch is 3–6, with the larger sizes after summer rains, and a nest may have as many as 12 eggs, apparently laid by two females (Bendire 1895, Pemberton 1925, Johnston 1964). Incubation period is 17–18 days (Bendire 1895, Sutton, in Bent 1940). Incubation starts when the second egg is laid, and early incubation before all eggs are laid results in asynchronous hatching and age differences of the nestlings in a brood. Both parents incubate, the male at night, both male and female in the daytime,

198 Lesser Roadrunner Geococcyx velox the female in one bout in the morning and one in the afternoon (Vehrencamp 1982). Both parents care for the young. Nestlings give a loud vocal “churr,” rattle the bill when disturbed and excrete a blackish foulsmelling liquid. Parents eat their nestlings’ fecal sacs and gain water this way (Calder 1968b).The parents feed the young nestlings with insects, and older nestlings are fed with lizards and small snakes while the adults feed themselves on insects. The weaker nestlings are eaten by the parent or are fed to stronger nestlings. Nestlings are 14 g at hatching, 50% of fledging weight by day 7, and fledge at 150 g, less than half the weight of the adults (Ohmart 1973, Folse 1974, Smith 1981).The young fledge in 17–19 days (Ohmart 1973). Natal down is lost on the day of fledging. The red and white mouth colors persist until day 50–55 when the mouth becomes marked with black, by day 80–85 the gape is black; the bare postorbital skin is orange

by 14 days (bluish color behind eye develops later). By 60 days the size and appearance of the young bird are similar to those of the adult (Muller 1971, Meinzer 1993). They forage actively with their parents when they are 70% of adult size, they follow their parents to feeding areas, and they catch their own food within 2–3 weeks (Sutton 1940). A parent sometimes flops a wing at its side and appears injured, a behavior that might draw a predator from the nest (Pemberton 1916).A pair of roadrunners in California renested a month after their first nesting, when the male took over care of the fledglings and the female laid a second set of eggs (Woods 1960). Nesting success: in Texas 73% of nests were taken by a predator, 22.5% of eggs survived to fledge (Folse and Arnold 1978), and a pair that nested twice could produce on average 3.5 chicks in a season. In Oklahoma and New Mexico, 66% of eggs laid hatched, 87% of nestlings fledged, and overall nest success was 72% (Maxon, in Hughes 1996b).

Lesser Roadrunner Geococcyx velox (Wagler, 1836) Cuculus velox A. Wagler, 1836, Gelehrte Anzeicher, München , 3, col. 96. (Mexico ⫽ outskirts of Mexico City) Monotypic.

Description ADULT: Sexes alike, crest and long tail, above the crown and nape black with white spots, short black crest with white spots, back chestnut brown with narrow white streaks, wing coverts chestnut brown with narrow white streaks, wing blackish with two white stripes across the primaries and the primaries with white tips, lower back and rump chestnut brown, tail long, blackish, with white tips on the outer rectrices T3–5 and white outline on the inner rectrices, underparts whitish buff, chin and upper throat white, sides of foreneck and breast whitish buff boldly streaked blackish, middle of neck and breast unstreaked, flanks and belly buff, under tail coverts sooty gray; bare skin behind eye blue to purplish, extending back to red patch on the neck, iris yellow to brown with ring of silvery white

around the pupil, bill above brown, darker on culmen, lower bill gray, tarsi gray to gray green. JUVENILE: Above buff tips to feathers, unstreaked, breast with dark spots not streaks, tail feathers narrower than in adult, bare skin around eye blue as in adult. NESTLING: Undescribed. SOURCES: AMNH, CM, CMNH, FMNH, MCZ, MVZ, ROM, UMMZ, USNM.

Measurements and weights Wing, M (n ⫽ 11) 133–154 (145.9 ⫾ 6.1), F (n ⫽ 9) 132–156 (142.1 ⫾ 8.5); tail, M 253–288 (271.8 ⫾ 11.6), F 247–280 (262.2 ⫾ 13.9); bill, M 34–40 (37.5 ⫾ 1.9), F 33–39 (36.9 ⫾ 2.3); tarsus, M 46–52 (48.7 ⫾ 1.8), F 43–50 (46.9 ⫾ 2.1) (UMMZ); Weight, M (n ⫽ 5) 174–203 (186.0), F (n ⫽ 4) 162.8–192 (174.0) (MVZ, UMMZ, USNM). Wing formula, P7 ⫽ 6 ⫽ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Lesser Roadrunner Geococcyx velox 199

History and variation The birds of Mexico were first collected by Europeans shortly after Mexican independence in 1821. One of the first collections was made by Fernando Deppe, a gardener and naturalist at Berlin University who left for Mexico in 1824.A roadrunner thought to be this species appears as a manuscript name “Geococcyx viaticus” in his specimen price list which was prepared in 1830 by his brother Wilhelm Deppe. F. Deppe sent specimens of these birds with a nest and two eggs to the Zoological Museum at Berlin, where Lichtenstein adopted the names of Deppe in his own collection register but did not publish or describe them (Stresemann 1954). In the Munich Museum, Wagler found another specimen taken by Deppe, and he described the species velox from Mexico and distinguished it from the G. californianus. Plumages vary with the season. Birds in fresh plumage in autumn are often rufous buff below, and the underparts become white by spring. Subspecies have been described (Moore 1934, Carriker and Meyer de Schauensee 1935, Peters 1940) based on color of the underparts, the pattern of white spots on the head, the pattern of white and black on the tip of the tail, and the shape of the tip of the bill. Because these traits also vary within a population in the tail, and with season in the tip of the bill and in color of the underparts, the following named subspecies are not recognized here: Geococcyx velox affinis Hartlaub 1844, arid subtropical zone of El Salvador and western Guatemala; G. v. longisignum Moore, 1934, Honduras and northern Nicaragua; G. v. melanchima Moore 1934, western Mexico from southern Sonora to Isthmus of Tehuantepec; G. c. pallidus Carriker and Meyer de Schauensee 1935, semi-arid lowlands of eastern Guatemala and Yucatan.

Field characters Overall length 48 cm. Large terrestrial cuckoo with long legs and long tail, cocked up or hanging or held straight behind as bird runs through the scrub. Lesser Roadrunner differs from Greater Roadrunner G. californianus in the bold streaks on side of throat and breast, the middle of the throat and breast unmarked, the rich buff color of the

sides and flanks, the chestnut not gray back and rump, and the dark under tail coverts.

Voice Call a series of soft “coo” notes, repeated on a descending scale, low-pitched around 0.5 kHz and often breaking into a lower voice at 0.3 kHz, each note in the series lower in pitch and voice breaking earlier in the note. Notes are about 0.5 sec long; the song has 5–7 notes in a series and lasts 5–6 sec.Also has a dove-like single “coo” at a higher pitch, about 0.6 kHz (Hardy et al. 1990, Howell and Webb 1995).

Range and status Western Mexico from southern Sonora to Oaxaca and Chiapas and on the central plateau from Nayarit and Jalisco south to México, Morelos, Puebla and Veracruz (local), and an isolated population occurs in Yucatán and extreme N Campeche. In Central America they occur from Guatemala and Honduras to El Salvador and central Nicaragua (Howell and Webb 1995, AOU 1998; UMMZ). Not threatened, although their habitat at lower elevations has been taken over by humans. Resident.

Habitat and general habits Arid lowland scrub, arid montane scrub and pine forest at higher altitudes; dry open country with scattered brush, thorn scrub, grassy and lightly wooded hillsides, farmland, habitat similar to that of Greater Roadrunner. In El Salvador the birds occur on upper slopes and in brushy gullies, above timberline on Volcán de Conchagua and Volcán de San Miguel and

200 Banded Ground-cuckoo Neomorphus radiolosus on the summit of Volcán de Santa Ana, there and on Volcán de San Salvador they also occur on the lower slopes below cultivation (Dickey and Van Rossem 1938). Sea level to 2800 m (Stotz et al. 1996, AOU 1998).

Food Insects, mainly grasshoppers and caterpillars (Dickey and Van Rossem 1938).

Breeding and life cycle In Mexico they nest in Veracruz in April, in Sonora in June and July, in Oaxaca from May to July, in Chiapas in April (Short 1974, Rowley 1984,

Binford 1989, Russell and Monson 1998; UMMZ 109131 from Chiapas); in Guatemala in April and May (Owen 1861), in El Salvador in August (Miller 1932, Dickey and Van Rossem 1938).The nest is a bulky stick platform, with a cup about 15 cm in diameter, built above ground in a bushy tree, a thorn bush or the center of an Opuntia cactus thicket. Eggs are white, 35 ⫻ 26 mm (Schönwetter 1964). Clutch 2, less often 3 or 4 (Owen 1861, Miller 1932, Dickey and Van Rossem 1938, Rowley 1984, Howell and Webb 1995).The incubation and nestling periods are unknown. Both male and female attend the nest (Dickey and Van Rossem 1938).

Genus Neomorphus Gloger, 1827 Neomorphus Gloger, 1827, in Froriep’s Notizen aus dem Gebiete der Natur-und Heilkunde, 16, col. 278. Large terrestrial cuckoos of New World tropical forests.Type, by original designation and monotypy, Coccyzus geoffroyi Temminck 1820. Spectacular birds

with bushy crests, striking plumage patterns and very long tails. The name Neomorphus refers to a new or different kind. The nostrils are broadly operculate and the opening slit-like, unlike in other cuckoos. Four species.

Banded Ground-cuckoo Neomorphus radiolosus Sclater and Salvin, 1878 Neomorphus radiolosus Sclater and Salvin, 1878. Proceedings of the Zoological Society of London 1878, p. 439, pl. 27 (Intac, Ecuador). Monotypic.

white, under tail coverts blackish; bare skin around eye blue, iris dark brown, bill deep, black to dusky above and paler below and near the tip, feet bluish gray.

Description

JUVENILE: Head dull black, back black and underparts black with rusty bars; bare skin around the face blue and gray (López-Lanús et al. 1999).

ADULT: Sexes alike, large terrestrial forest cuckoo, plumage mainly black, forehead glossy black with buffy white bars, crest glossy black, nape black, upper back black with narrow, buffy white bars, wing coverts dark purplish brown, inner primaries and secondaries deep purplish red with blackish inner vanes, outer primaries black, lower back brown with fine black bars, rump brownish black, tail long and black, central feathers glossed green to purple and outer rectrices glossed purplish; underparts blackish with scaly feathers barred buffy

NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH.

Measurements Wing, M (n ⫽ 2) 165–167 (166 ), F (n ⫽ 5) 162–172 (167.0 ⫾ 4.1), U (holotype) 177; tail, M 236–242 (239), F 236–254 (245.4 ⫾ 7.0), U 256; bill, M

Banded Ground-cuckoo Neomorphus radiolosus 201 45–46 (45.5), F 42–46 (44.8 ⫾ 1.6); tarsus, M 71–74 (72.5), F 68–74 (69.8 ⫾ 3.4) (AMNH, BMNH, FMNH). Wing formula, P5 ⱖ 4 ⱖ 3 ⬎ 2 ⬎ 1 ⬎ 6 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10; P1 to P5 are nearly equal in length.

Field characters Overall length 50 cm. Large terrestrial cuckoo of forests in its range, crest and hind neck glossy black, forehead barred white, underparts black with buffy white scallops, the black breast band indistinct and obscured by the blackish underparts, wings purplish red and black.

Voice Bill-snaps are given around ant swarms. Call is a deep-pitched “oooo” like that of a cow or a dove, repeated in an irregular pattern, with notes at 400 Hz.

Range and status South America in foothills of southwestern Colombia (lower Río San Juan southward) and northwestern Ecuador. In addition to museum records, they have been seen in Colombia (Valle, Cauca, Nariño, Río Ñambi valley) and Ecuador (Esmeraldas, Imbabura, Pichincha, Mache-Chindul hills). There were only three records in Colombia since 1956 until contact was made in with the birds

1997 in the Río Ñambi valley north of Junín, and only one in Ecuador since 1936 until the birds were observed in 1996–1998 at Estación Biología Jatun Sacha Bilsa in the Mache-Chindul hills, Esmeraldas, in northwest Ecuador (Hornbuckle 1997, Ridgely and Greenfield 2001). In Nariño they were reported to be fairly common in the 1990s. Resident. Rare and local, little known (Hilty and Brown 1986) and threatened due to widespread forest destruction (Collar et al. 1994).

Habitat and general habits Tropical lowland evergreen forest, in secondary forest patch within primary forest, in foothills and lower slopes on Pacific slope of the west Andes from 450 to 1525 m, with most records from 700 to 1200 m; in recent years in Ecuador recorded only below 500 m (Hilty and Brown 1986, Ridgely and Greenfield 2001). Individual birds forage on the ground and in the understory, often in mixed-species flocks at army ant swarms; the foraging flocks were mainly of Ocellated Antbird Phaenostictus mcleannani, also Plain-brown Woodcreeper Dendrocincla fuliginosa, Immaculate Antbird Myrmeciza immaculata, Bicolored Antbird Gymnopithys leucaspis and other ant-following birds. It perches on the ground or fallen logs, scours live leaves and stems of understory, and examines tree trunk bases from the ground. Moves rapidly on the ground at ant swarms, sprints in bursts and stops abruptly, runs forward to capture food, and when it catches prey items, it runs in a zig-zag pattern. It follows peccaries, and in Ecuador and Colombia it is known locally as “guide of the wild pigs” (AMNH 475939, 475942) or “seinero” (“running with the peccaries”); in Colombia it is the “correlona” (“fast runner”). The bare blue skin around the eye expands and contracts, perhaps in display (LópezLanús et al. 1999).

Breeding Unknown.

202 Rufous-winged Ground-cuckoo Neomorphus rufipennis

Rufous-winged Ground-cuckoo Neomorphus rufipennis (G. R. Gray, 1849) Cultrides rufipennis G. R. Gray, 1849. Proceedings of the Zoological Society of London, 1849, p. 63. [lower Orinoco River,Venezuela]. Monotypic.

Description ADULT: Sexes alike, forehead, crown and crest black glossed purplish, head and neck purplish blue black, back and wing coverts dark bronzy green, wing black primaries, secondaries dark brownish red, lower back to upper tail coverts dull olive, tail black with central feathers metallic purple, other tail feathers with a greenish gloss; underparts mostly unmarked, throat ashy white to gray, lower throat feathers gray with black borders forming a “V”, breast black, lower breast and belly dull gray brown to slate gray, under tail coverts slate gray; bare skin on face red to red-violet, skin on nape cobalt blue covered by black feathers, eye-lashes above and below the eye, iris brown, bill black with the tip greenish to greenish white, feet greenish gray to gray-blue. JUVENILE: Crown black, back and wing coverts brown-slate and loose-webbed, primaries black, secondaries dark brownish red as in adult, lower back and rump black, tail black, glossy purple and green, rectrices narrower than in the adult; underparts slate, feathers loose-webbed, skin black; facial skin dull red, skin behind ear cerulean blue covered by black feathers, eye-lashes above and below the eye, iris brown, bill black and not deep as in the adult. NESTLING: Hatchling, undescribed. A just-fledged young has black skin below and a few dark brown hair-like feathers attached to tips of the crown feathers. SOURCES: AMNH, ANSP, FMNH.

Measurements and weights Wing, M (n ⫽ 6) 164–176 (169.5 ⫾ 4.2), F (n ⫽ 6) 152–174 (164.3 ⫾ 8.3); tail, M 262–290 (275.7 ⫾

12.9), F 266–300 (288.5 ⫾ 14.5); bill, M 39–42 (40.7 ⫾ 1.4), F 38–44 (41.3 ⫾ 2.0); tarsus, M 63–71 (67.4 ⫾ 3.4), F 68–71 (69.8 ⫾ 1.6) (AMNH, ANSP, FMNH). Weight, M (n ⫽ 2) 350–520 (435); F (n ⫽ 2) 315–340 (327.5), and an emaciated fledged young F 144 (ANSP). Wing formula, P4 ⫽ 3 ⫽ 2 ⬎ 5 ⫽ 1 ⬎ 6 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 50 cm. Large ground cuckoo with bare red skin around the eye, above dark bronzy green, head, neck and breast glossy blackish and belly dull grayish brown.The red eye skin (red and blue in N. pucheranii) and breast pattern distinguish this species from other ground cuckoos.

Voice Single hoot “whoú”, a low call at 0.8–0.9 kHz and lasting 0.3 sec; also loud single snaps with the bill (Haffer 1977, Hardy et al. 1990).

Range and status South America in Amazonian Colombia (Chiribiquete: Stiles et al. 1995), southern Venezuela (Bolívar, northern and central Amazonas),

Red-billed Ground-cuckoo Neomorphus pucheranii 203 Guyana (Siparuni, Ituribisi, Supenaam and Mazaruni Rivers, Iwokrama Reserve near Kurupukari, Kamakusa, Annai and Bartica) and northern Brazil (Roraima near upper Rio Branco) (Chubb 1916, Snyder 1966, Haffer 1977, Meyer de Schauensee and Phelps 1978, Hilty and Brown 1986, ANSP). Resident. Rare and not well known.

Habitat and general habits Tropical lowland evergreen forest, seasonally flooded forest, also forest highlands in interior foothill zone, lowlands to ⬎1100 ? m. Solitary, wary, restless, terrestrial, runs fast on the ground, perches in the middle stratum of the forest canopy. Adults are seen on ant mounds (ANSP) and near bands of peccaries (Meyer de Schauensee and Phelps 1978).

Food Insects, including grasshoppers, crickets, spiders (Pelzeln 1871, ANSP).

Breeding An emaciated just- fledged short-tailed young was taken on the Siparuni River in Guyana on 17 March 1997, and a female had large ova 6 ⫻ 5 mm on 23 Sept 1997 (ANSP 187615, 188270).The nest is unknown. Eggs are yellowish white, 40 ⫻ 31 mm (Schönwetter 1964). Incubation and nestling periods are unknown.The fledgling is accompanied by the parents who care for it, as the just-fledged young was taken on the same ant mound where an adult male and female were collected a few days earlier.

Red-billed Ground-cuckoo Neomorphus pucheranii (Deville, 1851) Cultrides Pucheranii Deville, 1851, Revue et Magasin de Zoologie Pure et Appliquee (2), 3, p. 211. [Rio Yaguas, Peru] Polytypic.Two subspecies. Neomorphus pucheranii pucheranii (Deville, 1851); Neomorphus pucheranii lepidophanes Todd, 1925.

Description ADULT: Sexes alike, crest black glossed purple, back olive brown, rump brown, tail dark chestnut with central feathers T1 above olive glossed green, tail below appears black; underparts, throat unmarked gray, breast gray with blackish scales at tips of feathers, a black band on the breast, the band wider on flanks (15 mm) than in center (7–8 mm), belly buffy to light gray, flanks and under tail coverts dark brownish gray, wing purplish chestnut, outer primaries black glossed violet blue; face with bare red skin bordered behind with bright blue, iris brown to purplish red, bill red or orange with a yellow tip, feet dark gray. JUVENILE: Above brown without green gloss, crest blackish, wing purplish chestnut; underparts,

blackish throat and breast, belly brown, plumage loose-webbed, tail feathers narrower than in adult; facial skin bare, bill black, feet slate gray to black. NESTLING: Buffy hair-like natal down attached to tips of growing contour feathers on the head, fluffy plumage unmarked dark brown; bill black (BMNH 88.8.23.60). SOURCES: AMNH, BMNH, CM, LSU, ROM, UMMZ, USNM.

Subspecies Neomorphus pucheranii pucheranii (Deville, 1851); forehead light brown, breast feathers with narrow inconspicuous scaly tips, breast and belly buffy to light gray; Amazonian Peru and E Brazil north of River Amazon; Neomorphus pucheranii lepidophanes Todd, 1925; forehead lacks the distinct brown contrast with the black crown, breast feathers with conspicuous scaly tips, breast and belly light tan or clay color; Amazonian Peru and E Brazil south of River Amazon.

204 Rufous-vented Ground-cuckoo Neomorphus geoffroyi

Measurements and weights Neomorphus p. pucheranii: Wing, M (n ⫽ 6) 166–178 (171.7 ⫾ 4.8), F (n ⫽ 6) 160–168 (165.8 ⫾ 3.0); tail, M 266–285 (273.8 ⫾ 7.3), F 262–273 (267.0 ⫾ 3.6); bill, M 48–56 (51.9 ⫾ 3.2), F 44–55 (50.7 ⫾ 3.6); tarsus, M 65–68.5 (66.9 ⫾ 1.4), F 60.5–67.2 (65.9 ⫾ 1.1) (AMNH, CM, LSU, UMMZ); Neomorphus p. lepidophanes: Wing, M (n ⫽ 6) 162–182 (170.2 ⫾ 7.4), F (n ⫽ 5) 162–183 (169.8 ⫾ 8.0) (AMNH, CM, FMNH); Weight, M (n ⫽ 1) 330 (LSU). Wing formula, P5 ⫽ 4 ⫽ 3 ⫽ 2 ⬎ 1 ⬎ 6 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10; P1 to P6 nearly equal in length.

Field characters Overall length 50 cm. Large ground cuckoo, above dark bronzy green, head, neck and breast gray and belly dull gray, and red bare skin around eye and red bill.The red face differs from N. geoffroyi and N. radiolosus, and the red bill differs from N. rufipennis.

Voice The bird makes a guttural, roaring hum like a curassow, and it snaps the bill (Gyldenstolpe 1945a, Hilty and Brown 1986).

Range and status South America in Amazonian Peru (Río Curaray and Río Napo) and Brazil (Amazonia, from right bank of Rio Negro south to left bank of the Rio Purus) (Gyldenstolpe 1945a, Haffer 1977, Hilty and Brown 1986, Sick 1993). Sight observations have been reported in Amazonas, Colombia (Cuadros 1991), and in Amazonian Ecuador (Zábalo north of Rio Aguarico in E Sucumbíos,Yuturi in E Napo), the last

in near-sympatry with Rufous-vented Groundcuckoo N. geoffroyi (López-Lanús 1999, Ridgely and Greenfield 2001). Resident. Uncommon.

Habitat and general habits Tropical lowland evergreen forest, lowlands of upper Amazonia; lowlands to 700 m.Walks and runs on the ground, leaps from ground. It regularly forages at army ant swarms, and follows peccaries and tamarin monkeys Saguinus (Gyldenstolpe 1945a, Willis and Oniki 1978, Willis 1982, Siegel et al. 1989, López-Lanús 1999).

Food Insects, also plant matter (LSU).

Breeding In Peru a nestling was taken in February (BMNH). The nest and eggs are unknown except for a report of two eggs brooded by a pair (Castelnaux 1855, in Sick 1993).

Rufous-vented Ground-cuckoo Neomorphus geoffroyi (Temminck, 1820) Coccyzus geoffroyi, Temminck, 1820, Nouveau recueil de planches coloriées d’oiseaux, livre 2, pl. 7. [Pará, Brazil] Polytypic. Seven subspecies. Neomorphus geoffroyi geoffroyi (Temminck, 1820); Neomorphus geoffroyi salvini Sclater, 1866; Neomorphus geoffroyi aequatorialis Chapman, 1923; Neomorphus geoffroyi

squamiger Todd, 1925; Neomorphus geoffroyi dulcis Snethlage, 1927; Neomorphus geoffroyi maximiliani Pinto, 1962.

Description ADULT: Sexes alike, crest feathers black with a blue gloss, back and wing coverts bronzy brown, wing

Rufous-vented Ground-cuckoo Neomorphus geoffroyi 205 glossed green, outer primaries black with a purplish gloss, rump and upper tail coverts rufous brown, tail glossy black, forehead and underparts buff to cinnamon brown, breast feathers with dusky semicircular bands, breast with a narrow black band, lower breast buff, belly and under tail coverts rufous; bare facial skin blue, iris dark brown or with an inner ring of yellow and an outer ring of red, bill large and yellow-green to pale green, tarsi gray. JUVENILE: Plumage loose and fluffy, darker and duller than adult, head and back blackish, wing and tail black with greenish gloss, underparts blackish with rufous gray on belly; bare skin around eye gray with a bright blue spot behind the eye, iris dark brown to grayish yellow, bill black, tarsi bluish gray. NESTLING: Hatchling, undescribed. A just-fledged young bird in fluffy blackish plumage had a mouth lining of rose pink color, with small white papillae in the maxillo-palatine area. SOURCES: AMNH, BMNH, CM, FMNH, LSU, MCZ, MVZ, UCLA, UMMZ, USNM.

Subspecies Neomorphus geoffroyi salvini (Sclater, 1866); forehead and crown unbarred rufous brown, upperparts bronzy brown, breast feathers with broad pale tips, breast band distinct; Nicaragua (Caribbean slope), Costa Rica (mainly Caribbean slope, also Pacific drainage in Guanacaste), Panamá and the Pacific coast of Colombia; Neomorphus geoffroyi aequatorialis Chapman, 1923; breast feathers distinctly barred gray and buff; Ecuador, Peru, SE Colombia; Neomorphus geoffroyi geoffroyi (Temminck, 1820); forehead and crown brown barred with black feather edges giving a scaly appearance, upperparts olive green, breast with dusky bars formed with a dark brown subterminal bar on each buff feather, a lower breast band present; Brazil south of the Amazon, Pará; Neomorphus geoffroyi squamiger Todd, 1925; no black breast band; Brazil south of Amazon, lower Rio Tapajóz area;

Neomorphus geoffroyi australis Carriker, 1935; breast nearly unmarked in center and dusky bars on the side; Bolivia and Peru; Neomorphus geoffroyi maximiliani Pinto, 1962; upperparts dark bronze green, breast marks distinct; SE Brazil: Bahia; Neomorphus geoffroyi dulcis Snethlage, 1927; upperparts dark blue to green, breast marks distinct; SE Brazil from Espirito Santo to Rio de Janeiro, the range now a single locality in the Rio Doce basin. Seven subspecies are currently recognized in two groups, the first with the forehead and crown unbarred, the second with the forehead and crown barred. Genetic interchange between these populations in the not too remote past is suggested by intermediate populations between most of the extremes that are characterized in subspecies descriptions. Dispersal and genetic interchange between regions that are not separated by major rivers is suggested within Rufous-vented Ground Cuckoo N. geoffroyi, as (1) the color of the barred forehead and crown intergrade with salvini in NW Colombia having less rufous than salvini in Nicaragua (birds in Colombia are intermediate in color between salvini and aequatorialis), (2) the breast markings of the subspecies geoffroyi, squamiger and australis intergrade clinally, although locally in southern Amazonia and extreme northern Mato Grosso (Alta Floresta) the plumage is not well known, (3) the back color of geoffroyi and australis intergrade, and a Brazilian bird (AMNH 430469) taken in Rio Tocantins has a breast band that is nearly broken in the center and is intermediate between these two forms, and (4) the back colors of geoffroyi, maximiliani and dulcis vary clinally from green to blue (Griscom and Greenway 1941, Gyldenstolpe 1945a, Haffer 1977, Zimmer et al. 1997; AMNH, BMNH, LSU). N. g. squamiger is sometimes considered a separate species (Haffer 1977). However, the genetic distance between squamiger and N. geoffroyi salvini from Panama is much less than that between those birds and the other ground-cuckoos, whereas the genetic distances between the four species represented in Figure 5.3 indicate very nearly identical rates of genetic change within the Neomorphus groundcuckoos. Taken together, these genetic distances

206 Rufous-vented Ground-cuckoo Neomorphus geoffroyi indicate that squamiger is not a species distinct from N. geoffroyi.Also, examination of museum specimens shows intergradation across their geographic range between N. g. salvini and N. g. aequatorialis, and between N. g. geoffroyi, N. g. australis and N. g. squamiger. Early observers failed to find a nest of any Neomorphus species, and for a time there was speculation that these ground-cuckoos were broodparasitic. Sick (1949) discovered the parents and young together and established that these cuckoos cared for their offspring, and the adults have been seen with their nest.

bare colored skin on the face and a flat frontal erectile crest. Rufous-vented Ground-cuckoos are blackish or dark brown to rufous above, scaly brown on the breast, have a black breast band (lacking in squamiger) and rufous belly, and bare skin on the face gray to blue.

Measurements and weights

Range and status

Neomorphus geoffroyi salvini: Wing, M (n ⫽ 6) 165–175 (169.3), F (n ⫽ 11) 160–169 (163.6); tail, M 244–262 (254.8), F 237–266 (250.0); bill, M 43.6–49.5 (46.4), F 45.7–51.1 (47.9); tarsus, M 69.2–72.4 (71.6), F 69.5–73.1 (71.5) (Wetmore 1968); Nicaragua to Panamá, wing, M (n ⫽ 11) 167–184 (172.4 ⫾ 5.1), F (n ⫽ 15) 161–194 (170.1 ⫾ 7.9); tail, M 250–281 (263.3 ⫾ 10.7), F 247–278 (264.3 ⫾ 10.0); bill, M 40.0–52.3 (46.5 ⫾ 4.5), F 41–51.5 (43.3 ⫾ 3.0); tarsus, M 61–71 (67.6 ⫾ 3.2), F 63–71 (67.0 ⫾ 2.9) (AMNH, LSU, UMMZ, USNM); N. g. aequatorialis: M (n ⫽ 8) 162–188 (170.4 ⫾ 8.7), F (n ⫽ 2) 162–170 (166) (AMNH, FMNH); N. g. squamiger: M (n ⫽ 3) 163–171 (166), F (n ⫽ 2) 160–170 (165) (AMNH, CM); N. g. dulcis: M (n ⫽ 2) 172–173 (172.5), F (n ⫽ 1) 163, U (n ⫽ 2) wing 170–177 (173.2) (AMNH, BMNH). Weight, N. g. salvini: M (n ⫽ 1) 350, F (n ⫽ 2) 375–400 (387.5) (LSU); N. g. aequatorialis, M (n ⫽ 1) 340 (LSU); N. g. australis: Bolivia, M (n ⫽ 1) 370 (LSU); N. g. dulcis: M (n ⫽ 3) 339–355 (349), F (n ⫽ 1) 349 (Sick 1949, Haffer 1977); N. g. squamiger: M (n ⫽ 1) 340 (Graves and Zusi 1990). Wing formula, P6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Voice Loud bill-clapping, a soft dove-like moaning “oooooo-oóp”, a low “woof ” at army ant raids, and a loud “kchak!” in feeding competition (Slud 1964, Haffer 1977, Hilty and Brown 1986, Stiles and Skutch 1989, Hardy et al. 1990, Sick 1993).

Central and South America, from Nicaragua south through Costa Rica and Panamá to South America from western Colombia (Panamá border south on Pacific coast to Baudó Mts and east along N base of Andes in Córdoba to upper Río Nechí, and east of Andes in W Caquetá and W Putumayo), eastern Ecuador (close to base of Andes and from south of Río Napo, at Limoncocha, La Selva, Yuturi, Jatun Sacha, SE of Pompeya, Santiago in MoronaSantiago), Peru (except in eastern Amazonian region, where N. pucheranii occurs), NW Bolivia and Brazil (widespread except north of Rio Negro and Manaus in Rio Branco region where N. rufipennis occurs) (Howell 1957, Pinto 1962, Haffer 1977, Hilty and Brown 1986, Wetmore 1988, Remsen and Traylor 1989, Stiles and Skutch 1989, Dubs 1992, Sick 1993, Novaes and Cunha Lima 1998, López-Lanus 1999, Ridgely and Greenfield

?

Field characters Overall length 50 cm. Large terrestrial cuckoo of Neotropical forests, ground-cuckoos are fowl-sized, long-legged, long-tailed birds with a decurved bill,

Rufous-vented Ground-cuckoo Neomorphus geoffroyi 207 2001). Resident. Rare and local, solitary. They live in large blocks of natural forest, and like other ground-cuckoos they appear unable to cross the large rivers, which limit dispersal to movements along a river bank. May occur together with Redbilled Ground-cuckoo N. pucheranii near Yuturi in eastern Ecuador. This ground-cuckoo was one of the first birds to disappear from the protected but isolated forest at Barro Colorado, Panamá, where last seen in 1935 (Willis 1974, Eisenmann and Willis 1979). Population density, 0.25 pairs / 100 ha in floodplain forest in Amazonian Peru (Terborgh et al. 1990). Ground cuckoos in southern Brazil are disappearing due to widescale deforestation (Sick 1993). Because of its restricted range, N. g. squamiger are considered near-threatened (Collar et al. 1994). N. g. dulcis were formerly widespread in southeastern Brazil (Snethlage 1927, Pinto 1952) but have not been reported at the site where last observed, Rio Doce State Park, since before the 1980s, and these birds may be extinct (Knox and Walters 1994, Machado et al. 1998). Population density is sparse, at Cocha Cashu, Peru, estimated 0.25 territories in 100 ha of mature forest (Terborgh et al. 1990, Robinson and Terborgh 1997).

Habitat and general habits Tropical lowland forests, high ground forest and river bluff forest, and forest with seasonal flooding, in Ficus and Heliconia, canebrakes and thickets; in lowlands and foothills to 1000 m. Usually seen in riverine or slope forests in treefalls and vines, in dense undergrowth in primary forest and immature forest.Terrestrial, agile, can run, and can flutter to an elevated perch to lookout and to roost, but they are not seen in sustained flight. One or a pair (sometimes more, perhaps a family group) are seen at a time as they feed on forest floor in open undergrowth in mature forest. Feeding birds look into heaps of branches, armadillo holes and empty terrestrial termite mounds.Ant-following insectivores,

they forage at swarms of the army ants Eciton along with other ant-following birds, bills snapping at the edges of ant raids, tossing aside the leaf litter. They also follow peccaries. They leap over logs, dash about on the ground, flick aside leaves and peck at insects in the litter, and capture flushed out arthropods as the cuckoos take short runs along the ground (Sick 1949, 1953b, 1993, Howell 1957, Willis 1974, 1982, Terborgh et al. 1984, Stiles and Skutch 1989, Stotz et al. 1997). Crest erects when birds are excited.

Food Large insects, roaches, beetles; also scorpions, centipedes, spiders, small frogs, lizards, small birds, and occasional seeds and fruit (Pelzeln 1871, Sick 1949, Wetmore 1968, LSU).

Breeding In Central America breeds in the northern summer or wet season, in Nicaragua there was a fledged young in July, in Costa Rica there were fledged young in July and December. In northern Colombia a bird had an oviduct egg in April; in Mato Grosso, Brazil, they breed in September; in SE Brazil they breed in the southern summer (Sick 1949, Willis 1974, 1982, Roth 1981, Stiles and Skutch 1989; AMNH, BMNH, UMMZ, USNM). The nest is a broad shallow bowl of sticks, 25 cm across with a 12-cm cup, built 2–3 m above the ground in a shrub in dense swampy vegetation and lined with green leaves added through incubation. Eggs are yellowish white, 43 ⫻ 32 mm (oviduct egg, Wetmore 1968) to 40 ⫻ 32 mm (Roth 1981), clutch 1. The incubation and nestling periods are unknown.When disturbed, the incubating bird slips off the side of the nest and walks away. The first observation that ground-cuckoos rear their own young was made in 1941, when a pair was seen to attend their fledged young in Brazil (Sick 1949); another just-fledged young was taken with an adult male in Panamá (LSU).

208 Buff-headed Coucal Centropus milo

Centropodinae Genus Centropus Illiger, 1811 Centropus Illiger, 1811, Prodromus systematis mammalium et avium, p. 205. Large, heavy-bodied ground cuckoos of the Old World.Type, by subsequent designation, Cuculus aegyptius Gmelin (G. R. Gray, List Gen. Birds, 1840, p. 56). Large cuckoos with a stout bill, short rounded wings and long, broad tails. The genus name refers to the foot (Gr. kentron, a spike or spur, and pous, the foot), describing the long, straight hallux claw of most coucals; the German term for coucals “Sporenkuckucke” also refers to the spurred

foot. The broad wings cover the back; the exposed upper back (or mantle) is often the only region not covered by the wing; the back and rump are often covered by short blackish feathers that are concealed by the closed wing, in smaller species much as in the large broad-winged ground cuckoos (Carpococcyx, Neomorphus). The common name “coucal” may derive from “coucou” and “alouette”, the latter indicating the straight hallux claw as in a lark (“alouette”) (Cuvier, in Newton 1896). 26 species.

Buff-headed Coucal Centropus milo Gould, 1856 Centropus Milo Gould, 1856, Proceedings of the Zoological Society of London, 1856, p. 96, 136. (Guadalcanar, Solomon Islands) Polytypic. Two subspecies. Centropus milo milo Gould, 1856; Centropus milo albidiventris Rothschild, 1904.

Description ADULT: Sexes alike; buffy white head, upper back, throat and upper breast, lower back to rump and upper tail coverts black, wing black, tail long, graduated and black; belly either buffy white or black; iris red to red-brown, bill black, skin at base of bill dark gray, feet blue gray. JUVENILE: Upperparts and underparts brown mottled and barred with black, wing and long tail barred black and rufous; iris gray to brown, bill brown, lower mandible pale horn, whitish below, feet bluish gray. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, MVZ, UMMZ, UWBM.

Subspecies Centropus milo albidiventris Rothschild, 1904; breast, belly and thighs whitish; Solomon Islands (Vella Lavella, Bagga, Ganongo, Simbo, Gizo, Kulambangra, Kohinggo, New Georgia,Vangungu, Gatukai, Rendova,Tetipari); Centropus milo milo Gould, 1856; breast, belly and thighs black; Solomon Islands (Florida, Guadalcanal).

Measurements and weights Wing, M (n ⫽ 9) 260–283 (270.6 ⫾ 7.0), F (n ⫽ 9) 245–290 (266.8 ⫾ 16.5); tail, M 325–410 (358.0 ⫾ 25.0), F 312–372 (346.2 ⫾ 21.5); bill, M 52–63 (58.4 ⫾ 2.5), F 56–73 (63.9 ⫾ 5.8); tarsus, M 60–77 (65.6 ⫾ 6.4), F 67–74 (70.4 ⫾ 3.0); hallux claw, M 19–23.5 (21.3 ⫾ 1.8), F 17–23.5 (21.2 ⫾ 2.6) (AMNH, UMMZ, UWBM). Weight, M (n ⫽ 5) 742–790 (769.4) (UWBM). Wing formula, P 4 ⬎ 5 ⫽ 3 ⬎ 2 ⬎ 6 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 60–68 cm. Large coucal with upperparts, wing and tail black, with a buffy white

Buff-headed Coucal Centropus milo 209 head, upper back, throat and breast (and belly in some island regions). Juvenile is brown with black markings.

Voice A “sawing” or “snoring” noise, or a roar like a lion, “na-ow”’; members of a pair alternate calls “urrrh,” “-uh,” and a solitary bird gives a guttural “kkkkow . . . kkkk . . . kk . . . kk,” a throaty bark, and a deep churring “argh-a-argh,” and mellow grunts. Juveniles have a stuttering cough “koko, kokokkokoko . . .” (Sibley 1951, Cain and Galbraith 1956, Doughty et al. 1999, Diamond 2002, C. E. Filardi recording).

Range and status Solomon Islands (Vella Lavella. Gizo, Kuiambangra, New Georgia, Rendova, Guadalcanal and Florida island groups; absent on Choiseul, Ysabel, Malaita and San Cristobol island groups). Common resident (Hartert 1925b, Galbraith and Galbraith 1962, Blaber 1990, Sibley and Monroe 1990, Doughty et al. 1999, Mayr and Diamond 2001). Coucals apparently have had little success in dispersal between islands. In the New Georgia island group, coucals occurs on the nine largest islands, and on the 11th to 13th next largest islands, and are absent on smaller islands. In the Florida group they are

present on the largest island but not on the second largest, even though the islands are separated by only 150 to 1000 m of water. These birds are easy to detect because they roar loudly and they are well known by local people (Diamond 2002). The absence of large coucals on Buka and Bougainville may be the result of extinction after the Solomons were colonized from the Bismarcks by an ancestral coucal. extinction that might be owing to the human population with their introduced possums, rats, pigs and dogs. Or the absence may be due to another factor leading to extinction of coucals on the northern Solomons.

Habitat and general habits Lowland, primary and secondary forest and thickets, in undergrowth along stream in Casuarina hill forest, forested karst topography on coastal and raised limestone, hill and mist forest, Campernosperman plantations, and to 500 m on Kulambangra they occur around human habitation in villages and gardens, and secondary vegetation along the edge of airstrips. They live on the ground in the forest and outside it near the forest edge.They feed on the ground as they walk; and they are seen in trees and glide from tree to tree. They are mobbed by monarch Steel-blue Flycatcher Myiagra ferrocyanea, suggesting they take small birds and eggs and young in the nests (Sibley 1951, Blaber 1990, Doughty et al. 1999, C. Filardi, pers.comm.).

Food Insects, including stick insects, grasshoppers, beetles, pupae; giant centipedes (Sibley 1951, Blaber 1990, Doughty et al. 1999).

Breeding Males have enlarged testes (one per bird, or one with a vestigial second testis) in February, May and October (Sibley 1951, UWBM).The nest and eggs are unknown.

210 Pied Coucal Centropus ateralbus

Pied Coucal Centropus ateralbus Lesson, 1826 Centropus ateralbus Lesson, 1826, Bulletin Universel des Sciences et de l’Industrie, 8, sect. 2, p. 113. (New Ireland) Other common names:White-necked Coucal. Monotypic.

Description ADULT: Sexes alike, plumage black and white, and a long black tail. Plumage is variable in the extent of white. In the most common plumage phases, the cap and face are black, wing coverts black with a white patch, wing glossy purplish blue, neck and upper breast white, belly and under tail coverts black, and others are (1) all white with black forehead and lores, (2) all white on head and upper back, (3) black only on crown, (4) all black with white in wing patch, while other plumage variations have the white replaced by brownish white or pearly gray; iris red, bill black, legs slaty blue to black. Subadult or immature, blackish with white feathers on neck and wing. JUVENILE: Anterior half of body streaked and mottled with buff or rufous, rachi of hackles pale strawwhite, barbs lacking near tip, underparts plumage loose-webbed with barbs lacking; iris brown to gray, bill black, legs blue-gray. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, RMNH, ZMUC.

Measurements and weights Wing, M (n ⫽ 7) 192–217 (201.6 ⫾ 8.8), F (n ⫽ 7) 210–227 (214.1 ⫾ 10.4); tail, M 252–270 (263.0 ⫾ 8.8), F 195–227 (293.6 ⫾ 13.1); bill, M 39.5–44.0 (42.5 ⫾ 1.6), F 38.0–43.5 (41.2 ⫾ 1.9); tarsus, M 44.5–52.0 (49.4 ⫾ 2.9), F 45–54 (48.8 ⫾ 3.2); hallux claw, M 16–23 (19.3 ⫾ 2.3), F 18–28 (22.4 ⫾ 3.6) (AMNH, ZMUC).The hallux claw is large in the chick: 14 mm in a nestling with a tarsus of 37 mm, and 15 mm in one with a tarsus of 40 mm. Weight, M (n ⫽ 1) 330, F (n ⫽ 1) 342, U (n ⫽ 1) 345 (AMNH, USNM).

Wing formula, P6 ⬎ 7 ⫽ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 44–48 cm.A large coucal with black and white plumage, the only coucal like this in its range.

Voice Song a duet of two birds sounding like two hollow metal drums, “soo-hoo,” the duet lasting as long as 17 sec, the call a rising pitch, two birds drifting in and out of phase with each other. Alarm, a short “chit,”“krek” or “chunk”, sometimes repeated rapidly, also gives a nasal call “k-k-k-naaah!” (Coates 1985, 2001, Beehler et al. 1986). Dahl (1899) likened the call to the hollow toot of a village night watchman, and heard the call day and night.

Range and status Bismarck Archipelago (Umboi, New Britain, Lolobau, New Ireland, Dyaul); absent on the St. Matthais group and the western Bismarcks (Hartert 1925c, Gilliard and LeCroy 1967a, Orenstein 1976, Coates 1985, 2001, Mayr and Diamond 2001). Resident. The species was first seen by western eyes when Lesson, zoologist on the Coquille expedition, visited the Bismarck Archipelago in 1823; and it was one of the first bird species recorded by the first European explorers to the region. Lesson described

Greater Black Coucal Centropus menbeki 211 the bird as the “coucal atralbin” (Mayr and Diamond 2001).

Habitat and general habits Forest, forest edge, secondary forest, vegetation on sea shore, coconut plantations and gardens.They mainly frequent shrubbery and regrowth. Occur in lowlands near the coast and in hills, locally to 1200 m in New Britain and 1400 m in New Ireland, more common in New Britain than in New Ireland.They climb in coconut trees, walking on the midribs of hanging fronds, gleaning in and around the coconuts and in the center of a tree, and in the undergrowth in overgrown coconut plantations (LeCroy and Peckover 1983).They forage from the ground climbing to the tops of shrubs and into high trees, gliding down to a low level again, active and clumsy in movements. They occur in pairs and in family groups numbering up to four birds (Hartert 1926, Coates 2001).

Food Insects, including longicorn beetles Cerambycidae and large stick insects Phasmatodea, lizards, probably other small animals (Dahl 1899, Coates 1985, 2001).

Breeding Feathered nestlings and recently fledged young are known from November to February and a justhatched nestling was seen in February (Dahl 1899). Feathered nestlings have also been taken in May (ZMUC, Noona Expedition, 2.1.1982.4) and in July or August and (Rothschild and Hartert 1914). Nest is a hollow or chamber, built of dry stems and leaves of reed-grasses, sometimes with two holes front and back, lined with green vegetation. Eggs are whitish, 41 ⫻ 34 mm, clutch 2 or 3 (Meyer 1933, 1936, Schönwetter 1964). The incubation and nestling periods are unknown.

Greater Black Coucal Centropus menbeki Lesson and Garnot, 1828 Centropus Menbeki Lesson and Garnot, 1828, Voyage autour du monde sur la corvette la ‘Coquille,’ Atlas, pl. 33; Zoologie, 1, livre 13, 1829, p. 600. (New Guinea (Dorey, ⫽ Manokwari)) Other common names: Greater Coucal, Menbek’s Coucal, Black Jungle Coucal. Polytypic. Two subspecies. Centropus menbeki menbeki Lesson and Garnot, 1828; Centropus menbeki aruensis (Salvadori, 1878).

JUVENILE: Upper-parts and underparts dull blackish, some neck and upper back feathers with strawcolored rachi, wing black sometimes with faint rufous bars, tail black with indistinct narrow rufous bars, the feathers 40–50 mm broad at mid-length and nearly pointed at the tip; iris light gray (AMNH 425917, 425920, 766255).

Description

SOURCES: AMNH, ANSP, BMNH, FMNH, MCZ, MSNG, MVZ, MZB, RMNH, ROM, SMTD, USNM, ZFMK.

ADULT: Sexes alike, huge size, upperparts and underparts glossy black, stiff black hackles on forepart of body, wing black, tail long, graduated, glossy black, the feathers 60–70 mm broad at midlength and wide at the tip; iris red, bill whitish with black base, legs black. SUBADULT: (or NON-BREEDING): Upperparts dull black, stiff black hackles, tail black with distinct narrow rufous bars on all but the tip, the feathers 50–60 mm broad at mid-length and rounded at the tip (AMNH 425915).

NESTLING: Undescribed.

Subspecies Centropus menbeki menbeki Lesson and Garnot, 1828; plumage glossed blue to blue-green; New Guinea and the western Papuan islands,Yapen Is; Centropus menbeki aruensis (Salvadori, 1878); plumage glossed purple; Aru Islands. History: An earlier account and illustration by Martin in 1785, “The Green Bird of Paradise” had a red iris, glossy black plumage with scale-like body

212 Greater Black Coucal Centropus menbeki feathers, a long tail, zygodactyl feet and a short hallux claw, but had a black bill (Stonor 1937). Geographic variation: Birds of Yapen (Japen, Jobi) are more blue (less green) than certain birds of New Guinea, and were described as a subspecies C. m. jobiensis Stresemann and Paludan 1932, but birds with the same blue plumage color occur on the Sepik River (Rothschild et al. 1932), the Fly River area (AMNH, Rothschild and Hartert 1907) and on Misool (ANSP, Mayr and Meyer de Schauensee 1940c).

Measurements and weights C. m. menbeki, New Guinea: Wing, M (n ⫽ 13) 210–240 (224.5 ⫾ 9.1), F (n ⫽ 10) 209–240 (220.8 ⫾ 8.3); tail, M 344–411 (367.9 ⫾ 23.0), F 326–391 (355.2 ⫾ 17.5); bill, M 46–54 (49.1 ⫾ 3.5), F 44–55 (50.3 ⫾ 3.8); tarsus, M 55–64 (59.0 ⫾ 3.2), F 57–61 (59.6 ⫾ 1.4); hallux claw, M 16–23 (20.3 ⫾ 2.9), F 18–21 (19.6 ⫾ 1.1) (AMNH, FMNH, MSNG, RMNH);Yapen:Wing, M (n ⫽ 3) 225–235 (230.7), F (n ⫽ 2); 224–225 (224.5); tail, M 350–360 (355), F 333–360 (346.56) (AMNH, FMNH, MSNG); C. m. aruensis, Aru Islands: Wing, M (n ⫽ 2) 234–238 (236), F (n ⫽ 1) 260; tail, M 328–378 (353), F 416 (AMNH, SMTD). Weight, New Guinea: M (n ⫽ 6) 430–575 (493), F (n ⫽ 2) 505–553 (529) (AMNH, FMNH, MZB, RMNH, Hartert 1930, Ripley 1964);Yapen: F (n ⫽ 1) 450 (FMNH). Wing formula, P4 ⫽ 3 ⫽ 2 ⬎ 1 ⱖ 5 ⬎ 6 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 67 cm. A large black coucal of the forests of New Guinea and the Aru Islands.

Voice Low-pitched, resonant booming hoots staccato, single note “oodle,” also pairs of “hoo hoo,” or four hollow “hoo” notes, and a series that descends in pitch “Uh-oo-oo-oo-oo-oo-oh”, a grunt or croak followed by a dry rattle.The call is like “an un-oiled door being shut” (Gyldenstolpe 1955). Calls are deeper and delivered more slowly than in Pheasant Coucal C. phasianinus in New Guinea and more complex and longer than in Lesser Black Coucal C. bernsteini. Pairs call in duet. Coucals sometimes

call after sunset and at night (Rand 1942a, Diamond 1972, Coates 1985, Beehler et al. 1986).

Range and status New Guinea, western Papuan islands (Batanta, Salawati, Misool), Yapen and Numfor islands, and Aru Islands (Salvadori 1881, Rothschild and Hartert 1907, Hartert 1930, Rothschild et al. 1932, Hartert et al. 1936, Maxwell 1938, Mayr and Meyer de Schauensee 1940b, Mees 1982, Gregory 1995, Coates 1985, Sibley and Monroe 1990). Resident. Widely distributed, uncommon. Density, c. 1 bird / 10 ha (Bell 1982a).

Habitat and general habits Forest, forest edge, in low undergrowth, shrub and lower middle stories, and in swampy grasslands; from sea level to 850 m (Rand 1942a,b), while in the eastern highlands of New Guinea, they have been seen at Karimui at 1275 m, well above their usual altitudinal range (Diamond 1972). Feed on the ground, also hop up tree trunks and vines while they switch the tail from side to side. Terrestrial, they seldom fly (Ogilvie-Grant 1915, Stein 1936, Diamond 1972, Coates 1985, Beehler et al. 1986).

Food Small vertebrates (snakes, frogs, small birds), arthropods, large insects (grasshoppers, cicadas, caterpillars, beetles) taken on the ground (Rothschild et al. 1932, Rand 1942a,b).

Breeding Near the Idenburg River in northern West Papua a female had an oviduct egg in April, fledged young

Biak Coucal Centropus chalybeus 213 were seen in the middle Sepik River area in January and near the Fly River in October, and a female was in breeding condition in September (Rand 1942a, b, Gilliard and LeCroy 1966). Nest is a large mass of

leaves, in pandanus, built in the wet season.The egg, known from the oviduct of a breeding female, is white, oval, 37 ⫻ 30.3 mm (Rand 1942b). Clutch size, incubation and nestling periods are unknown.

Biak Coucal Centropus chalybeus (Salvadori, 1875) Nesocentor chalybeus Salvadori, 1875, Annali del Museo Civico di Storia Naturale di Genova, 7, p. 915. (Misori ⫽ Biak) Monotypic.

Description ADULT: Sexes alike, plumage black, upperparts with a dull purple gloss, foreparts with spiny feathers, wing black, long black tail, underparts black to dark brown on belly; iris yellow, bill black, legs and feet black. JUVENILE: Plumage unbarred, dark and duller than in adult, some birds are washed rufous.

Voice NESTLING: Undescribed. SOURCES: AMNH, ANSP, MSNG, MZB, SMTD, ZMB.

Measurements

Loud, hollow notes, a series of upslurred “hoot” notes, sometimes slightly accelerated, the notes descending or going up and down in waves, also a harsh rasp, and a repeated “bup”; a noisy bird (Beehler et al. 1986).

Wing, M (n ⫽ 5) 189–214 (203.0), F (n ⫽ 6) 189–220 (203.0 ⫾ 9.7); tail, M 260–285 (274.2), F 253–315 (280.3 ⫾ 23.0); bill, M 44–50 (47.2), F 40–49.3 (45.0 ⫾ 3.1); tarsus, M 47–54 (50.4), F 50–57 (53.6 ⫾ 2.4); hallux claw, M 15.0–16.2 (15.6), F 15.2–16.2 (15.6) (AMNH,ANSP, MSNG, SMTD, ZMB). Wing formula, P5 ⬎ 4 ⬎ 3 ⬎ 2 ⫽ 6 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10 (P1 is half the length of P5).

Range and status

Field characters

Habitat and general habits

Overall length 44–46 cm. Small coucal with black plumage, a relatively short tail and a yellow iris.This is the only coucal on Biak Island. It is smaller and has a straighter bill than the Greater Black Coucal C. menbeki of New Guinea, and a stockier build and yellow iris unlike the Lesser Black Coucal C. bernsteini of New Guinea.

Primarily lowland forest, also thick second growth, on ground, in vines and in trees.Terrestrial, it hops and feeds on the ground. It is awkward in flight, like other coucals.

Biak Island, in Geelvink Bay north of Irian Jaya (Salvadori 1875, Mayr and Meyer de Schauensee 1940a, Rand and Gilliard 1967, Beehler et al. 1986). Resident. Hunting and habitat destruction have reduced the numbers in southern Biak, and the main population is in the forests of Supiori (Bishop 1982); conservation status near-threatened (Collar et al. 1994).

Breeding Unknown.

214 Rufous Coucal Centropus unirufus

Rufous Coucal Centropus unirufus (Cabanis and Heine, 1863) Pyrrhocentor unirufus Cabanis and Heine, 1863, Museum Heineanum, 4(1), p. 118. (Philippines [Luzon]) Monotypic.

Description ADULT: Sexes alike, plumage rufous chestnut, head slightly darker, wing and tail uniform rufous chestnut; bare skin around eye yellow, iris light brown, bill green with yellow tip, feet black. JUVENILE: Like adult, rectrices narrow; bill black. NESTLING: Yellowish-white hair-like down on head. SOURCES: AMNH, CM, FMNH, USNM.

Measurements and weights Wing, M (n ⫽ 11) 145–162 (155.7 ⫾ 5.6), F (n ⫽ 8) 155–165 (161.5 ⫾ 3.3); tail, M 213–242 (228.2 ⫾ 1.9), F 220–262 (237.6 ⫾ 12.8); bill, M 32–36 (33.8 ⫾ 1.1), F 33–38 (35.3 ⫾ 1.6); tarsus, M 36–45 (40.2 ⫾ 3.3), F 39–45 (41.8 ⫾ 2.5); hallux claw, M 16–20 (18.2 ⫾ 1.5), F 17–19 (17.9 ⫾ 0.6) (FMNH). Weight, M (n ⫽ 6) 145.7–187.7 (168.3), F (n ⫽ 8) 146–227.9 (201.4) (FMNH, USNM, Goodman and Gonzales 1990). Wing formula, P5 ⫽ 4 ⬎ 3 ⬎ 6 ⬎ 2 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 38–42 cm. Coucal with uniform light brown to rufous plumage, pale iris and green bill.

Voice Call, a snapping staccato of two to five pipsqueak notes,“squip-whip” or “squip-whip-whip . . .”, high in pitch at 4–5 kHz as the birds move through the undergrowth; also squeaks and short metallic trills, and a shrill, plaintive “kaow” when a bird is separated from its mate (Gilliard 1950, Scharringa 1999, Kennedy et al. 2000).

Range and status N Philippines (Luzon, Polillo, Catanduanes). Resident. Uncommon and local (Stresemann 1939, Gilliard 1950, Gonzales 1983, Dickinson et al. 1991). Because of its restricted range and the loss of forest habitat in the Philippines, the species is considered near-threatened (BirdLife International 2001).

Habitat and general habits Hill forest in matted undergrowth, tangled lowland forest with bamboo, usually on or near the ground, also in trees. They live in the understory in dense undergrowth and are associated with climbing vines and bamboo thickets. Occur both in primary forest and in selectively logged forest (Poulsen 1995). On Mt Isarog they were seen between 550 and 760 m in 1961, but they were not seen there in 1988 (Goodman and Gonzales 1990). The behavior as it moves about is more like a malkoha than a typical coucal. Seen singly, in pairs and in groups of up to 10 to 12 birds, noisy and walking away from disturbance; once seen following a mixedspecies flock of birds (Poulsen 1995, Kennedy et al. 2000).

Food Unknown.

Green-billed Coucal Centropus chlororhynchos 215

Breeding A female had an egg ready to lay on 14 April, and a juvenile with a short wing and tail and with hair-

like down on the head in late November is from an October nesting (FMNH). The nest and eggs are unknown.

Green-billed Coucal Centropus chlororhynchos Blyth, 1849 Centropus chlororhynchos Blyth, 1849, Journal of the Asiatic Society of Bengal 18, pt. 2, p. 805. (Ceylon) Note: Often spelled Centropus chlororhynchus, the original description is Centropus chlororhynchos. Other common name: Ceylon Coucal, Sri Lanka Green-billed Coucal Monotypic.

Field characters

Description

Voice

ADULT: Sexes alike, upperparts and underparts dull black, neck and tail glossed purple, the back, wing coverts and wing chestnut in no great contrast to the upperparts, tail black, wing lining black; iris red to reddish brown, bill ivory to pale green, legs and feet black.

Song is a resonant, sonorous, low-pitched two or three notes “hoop-poop-poop,” the last note lower in pitch, the song 0.3–0.35 kHz, given at 2 notes per sec, the calls often given as a duet; also a short double syllable “hu, hu,” or “hoooooEE” and a single cough “chewkk” (Hoffmann 1989a,b, Grimmett et al. 1999,Wijesinghe 1999).

JUVENILE: Plumage like adult’s, but wing coverts indistinctly barred black; iris gray, bill dark greenish with base and culmen gray, legs and feet dusky gray. NESTLING: Skin black, white hair-like down on head and neck, bill with an ivory-colored stripe on bill from lores to tip of upper mandible, lower mandible black (Wijesinghe 1999). SOURCES: ANSP, BMNH, MCZ,YPM, ZMB.

Measurements Wing, M (n ⫽ 7) 158–173 (167.3 ⫾ 4.9), F (n ⫽ 7) 168–184 (174.6 ⫾ 7.7); tail, M 222–252 (234.3 ⫾ 8.8), F 225–254 (239.3 ⫾ 10.7); bill, M 36–41 (38.9 ⫾ 1.7), F 38.9–47.0 (42.0 ⫾ 2.9); tarsus, M 40–46 (43.3 ⫾ 2.1), F 42–47 (42.9 ⫾ 2.6), hallux claw, M 15–18 (17.1 ⫾ 1.1), F 17–21 (18.1 ⫾ 1.5) (BMNH). Wing formula, P6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 2 ⬎ 7 ⬎ 8 ⬎ 1 ⬎ 9 ⬎ 10.

Overall length 43–46 cm. Dull blackish head, body and tail, and dark rufous wings. On Sri Lanka, distinguished from the larger Greater Coucal C. sinensis by pale green rather than black bill, by its dark wings, and by its shorter deep call lacking the water-bottle phrase; occurs in wet forest habitat.

Range and status Sri Lanka. It occurs from the center (Dambulla to Ambepussa and Kandy) to the southwest. Observations since 1980 are from Amanawala Ampane, Kitulgala, Htaramune and Dehiowita to Morapitiya Forest Reserve, Neluktiya Mukalana,

216 Black-faced Coucal Centropus melanops Ingiya Forest Reserve, Peak Wilderness Area, Gilimale forest, Sinharaja Forest Reserve, and Nakiyadeniya and other sites in this region (BirdLife International 2001). Resident. Local forests are separated by extensive areas of cultivation and coffee plantations, habitats that may discourage dispersal between forest sites.The total population is perhaps a few thousand individuals. The most securely protected forest site that supports these coucals is on the lower edge of Peak Wilderness Sanctuary in Sinharaja National Heritage Wilderness Area, Ratnapura District. Globally threatened, endangered by loss and fragmentation of forest habitat as the coucals’ home is converted into agricultural land and the remaining forests are cut and cleared of undergrowth (BirdLife International 2001).

Food Omnivorous, they take beetles, grasshoppers, termites, spiders; snails, ammonites, worms; frogs, small lizards including skinks, and snakes as long as 60 cm, and they also feed on plant material (Legge 1880, Hoffmann 1989a,b,Wijesinghe 1999).

Breeding behavior Courtship behavior has not been described, nor has courtship feeding, although the male brings food to the nest while the female incubates, and the male is said to remove her fecal sacs from the nest. The adult bill color changes from whitish to green through the nesting period (Wijesinghe 1999).

Breeding Habitat and general habits Humid high evergreen forest with dense undergrowth, dense scrub with bamboo and rattan cane rushes in disturbed areas, tangled thickets, wooded river banks, occur in wet zone forests west, southwest and south of the central mountain massif; low country to 760 m (Legge 1880, Lewis 1898, Wait 1925, BirdLife International 2001). The coucal is closely associated with the bamboo Ochlandra stridula, which grows in wet places, and the Sinhala name for this bamboo is bata aetikukula, or “coucal bamboo” (Wijesinghe 1999). Occasionally appears in patches of abandoned slash-and-burn agriculture close to forest and in other disturbed areas with tangled vegetation (BirdLife International 2001). The coucals feed on the ground and in trees and creepers. In the morning they perch in the open and spread their wings to the sun (Wijesinghe 1999).

Prolonged season with nesting records in January, April to July, September and November. Nest is a covered mass of twigs, roots and grass, with a side entrance, lined with green or dried leaves especially bamboo. The nest is built in thorn canes, bamboo or in a tree, near the ground to as high as 3.6 m above ground in a tree, especially in Wendlandia bicuspidata (Rubiaceae). Both sexes build the nest, which measures 60 c m ⫻ 45 cm. Eggs are chalky white, 35 ⫻ 27 mm, the clutch is 2(3). Incubation period 17 days or shorter; nestling period 12 days or longer (hatching day was not determined, Wijesinghe 1999). Both parents feed the young. The young fledgling flutters the wings and tail when it is fed. It returns to roost in the nest for the first two days after it leaves the nest (Wait 1925, Baker 1934, Wijesinghe 1999, BirdLife International 2001).

Black-faced Coucal Centropus melanops Lesson, 1830 Centropus melanops Lesson, 1830, Traité d’Ornithologie, livre 2, p. 137. (“Java” [⫽ Mindanao]) Monotypic.

Description ADULT: Sexes alike, center of crown buffy white, neck darker pinkish buff, upper back rufous, wing

rufous with dark tips to primaries and secondaries, lower back and upper tail coverts black, tail black, face whitish with a black mask from forehead around eye to base of lower mandible; underparts, chin and throat buffy white, breast buff, lower breast and belly to under tail coverts black; iris red, bill black, feet black.

Black-faced Coucal Centropus melanops 217 JUVENILE: Like adult, rectrices narrow. NESTLING: Undescribed. SOURCES: AMNH, CM, CMNH, FMNH, MVZ, RMNH, ROM, UMMZ, ZMUC.

Geographic variation A subspecies C. m. banken Hachisuka 1934 was described from Bohol, Leyte and Samar, with more black on the face, but these do not differ consistently between birds on these islands and birds elsewhere in the Philippines (Rand and Rabor 1960, Parkes 1971; AMNH, FMNH).

Measurements and weights Wing, Mindanao: M (n ⫽ 8) 152–167 (162.0 ⫾ 4.1), F (n ⫽ 5) 158–172 (162.0 ⫾ 5.4); tail, M 222–252 (237.7 ⫾ 10.6), F 215–270 (243.8 ⫾ 20.2); bill, M 30–37 (34.6 ⫾ 2.9), F 33–36 (34.1 7 ⫾ 1.2); tarsus, M 34–47 (42.0 ⫾ 5.5), F 39–44 (41.3 ⫾ 1.7), hallux claw, M 15.5–18 (16.7 ⫾ 0.9), F 14–17 (16.0 ⫾ 1.2) (CMNH, ZMUC); Weight, M (n ⫽ 10) 197.5–237.3 (211.1), F (n ⫽ 12) 214–265.5 (237.9) (Rand and Rabor 1960; AMNH, CMNH, USNM). Wing formula, P6 ⫽ 5 ⫽ 4 ⬎ 7 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 42–48 cm. Coucal with buff foreparts and a black face with whitish crown streak running forward to the bill, rufous back and wings and a black tail.

Voice Call is a distinctive loud booming “wooop wooop wooop wooop wooop”, the first note longer and lasting 0.5 sec, then a pause of a second, the

other notes becoming softer, the series lasting about 5 sec. Other calls are a single “wooop” and a descending “boop boop boop boop” (Kennedy et al. 2000).

Range and status Philippines (Basilan, Mindanao, Nipa, Dinagat, Sirgao, Bohol, Leyte, Samar). Resident. Common in Mindanao 60 years ago, its forests are nearly gone (Hachisuka 1934, Dickinson et al. 1991).

Habitat and general habits Forests, in treetops, second growth; lowlands to 1200 m.

Food Unknown.

Breeding April (oviduct egg).The nest is undescribed. Eggs are white, fine texture, not glossy, 31 ⫻ 26 mm (Rand and Rabor 1960) or 37 ⫻ 29 mm (Schönwetter 1964). Incubation and nestling periods are unknown.

218 Black-hooded Coucal Centropus steerii

Black-hooded Coucal Centropus steerii Bourns and Worcester, 1894 Centropus steerii Bourns and Worcester, 1894, Occasional Papers of the Minnesota Academy of Natural Sciences, 1, p. 14. (Mindoro) Other common names: Steere’s Coucal. Monotypic.

Voice

Description

Range and status

ADULT: Sexes alike, head black glossed blue green, hackles on head and throat black, neck, back and wing sooty brownish black, tail glossy greenish black, chin and throat black, breast and belly dark brown; iris red to brown, bill black, legs and feet black.

Philippines (Mindoro) (Dickinson et al. 1991). Resident. Rare and near extinction as a result of loss of their lowland primary forest. Where forest habitat has been degraded, these coucals are replaced by the Philippine Coucal C. viridis. Blackhooded Coucals are known from recent observations in four sites: Puerto Galera, MUFRC Experimental Forest (the only inland site), Malpalon (where rough ground has given the forest respite from clearing and cultivation) and Siburan (where the Sablayan Penal Colony has provided restricted human access into the forest, although prisoners now disturb the forest undergrowth collecting rattan and logging for furniture production; and refugees who were resettled to Mindoro from their homes in Luzon in 1991–1993 during a period of volcanic activity on Mt Pinatubo have had an impact on the forest as well). The conservation status is critical, as this coucal is rare and lives in very small fragmented forest sites, which are few and mostly not effectively protected. It is the next candidate cuckoo for global

JUVENILE: Plumage above brown, soft-webbed and lacking hackles, tail dull black, chin and throat brownish gray; bill pale and depth less than in adult. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, CM, FMNH, SMF, YPM.

Measurements and weights Wing, M (n ⫽ 8) 152–171 (159.4 ⫾ 5.7), F (n ⫽ 9) 157–170 (164.4 ⫾ 5.0); tail, M 217–253 (238.6 ⫾ 11.5), F 232–277 (251.3 ⫾ 16.6); bill, M 35–40 (36.9 ⫾ 1.9), F 35–41 (37.1 ⫾ 1.9); tarsus, M 35–45 (39.6 ⫾ 3.4), F 38–45 (43.2 ⫾ 3.7); hallux claw, M 16–17 (16.5 ⫾ 0.5), F 16–18 (16.8 ⫾ 1.2) (AMNH, BMNH,YPM, SMF). Weight, M (n ⫽ 4) 160–200 (179), F (n ⫽ 7) 190–238 (163.4) (BMNH, YPM, Ripley and Rabor 1958. Wing formula, P5 ⫽ 4 ⬎ 3 ⬎ 7 ⫽ 6 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 46 cm. An all dark coucal with head black, back and wing fuscous, and a glossy greenish black tail. Larger and stockier than the black form of Philippine Coucal C. viridis which also lives on Mindoro; plumage sooty brown not glossy black.

Resonant, deep booming notes, “hoot hoot hoot hoot . . .”, 5 to 9 or more notes in a series, the first note louder and longer, the final notes descending slightly and becoming softer (Kennedy et al. 2000).

Short-toed Coucal Centropus rectunguis 219 extinction (Dutson et al. 1992, Brooks et al. 1995, BirdLife International 2001).

Habitat and general habits Primary old growth dipterocarp forest in lowlands, and in mountains to 800 m in transition dipterocarp-mid-mountain forest. The birds live in primary forest on rocky ridges with small forest openings with bamboo and creepers, in tangled thickets of vines and rattan, and streamsides within primary forest, all minimally disturbed forest habitats (Bourns and Worcester 1894, McGregor 1905,

Kennedy et al. 2000, BirdLife International 2001). Secretive, they move slowly through dense vines and foliage, in understory and forest canopy; they run along tree branches and spend time in thick tangles of forest vines and creepers.

Food Unknown, save for a specimen noted to have “larger insects” (BMNH).

Breeding Unknown.

Short-toed Coucal Centropus rectunguis Strickland, 1847 Centropus rectunguis, Strickland, 1847, Proceedings of the Zoological Society of London, 1846, p. 104. (Malacca)

Description ADULT: Sexes alike, above especially the nape glossy purplish-blue black, shafts of hackles black, wing chestnut, tail black, underparts black, under wing coverts black; iris red, bill black, legs and feet black. Hallux claw short and straight, not long as in most coucals. No eclipse plumage. JUVENILE: Head above chestnut rufous with brown shaft streaks, back rufous with indistinct blackish bars, wing coverts chestnut with black bars, flight feathers unbarred chestnut with no dark on the edges or tips, tail black with fine white bars; chin to belly dark brown barred with dull white, V-shaped bars on each feather of throat and breast and fine wavy bars on the belly (some birds are nearly black below, others are brown or buff ), under wing coverts black; iris gray, bill blackish horn to brown, paler below, legs and feet blackish. NESTLING: Undescribed. SOURCES: AMNH, BMNH, ZRC.

Measurements and weights Wing, M (n ⫽ 4) 156–170 (165.5), F (n ⫽ 2) 166–186; tail, M 192–204 (198), F 194–238 (216);

bill, M 35–37 (36), F 34–42 (37); tarsus, M 42–46 (44.3), F 45–52 (48.5); hallux claw, M 8–12 (10.0), F 9–19 (14) (AMNH, BMNH, ZRC). Weight, F? (a large unsexed bird) (n ⫽ 1) 237.5 (Wells 1999). Wing formula, P6 ⫽ 5 ⫽ 4 ⫽ 3 ⬎ 2 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 38–43 cm. Black plumage with uniformly rufous wing and black tail, wing and tail shorter than Greater Coucal C. sinensis, and hallux claw short and nearly straight.

Voice Call, deep resonant booming notes on a descending scale, “whu, huup-huup-huup-huup” at a pitch of 0.3 kHz, given 2 notes per sec, each note lasting 0.25 sec; and a rapid series of deep rising notes heard at dusk. A snorting squirrel-like “hut, hut, hut” and an explosive scold “jeézaw” may also be from this coucal (Smythies 1981, Scharringa 1999, Wells 1999, Robson 2000a, Sheldon et al. 2001).

Range and status Malay Peninsula, Sumatra and Borneo. Exact distribution is not well known and many published records are questionable, especially reports in upland habitats where observations may be of the lookalike Larger Coucal Centropus sinensis (BirdLife

220 Short-toed Coucal Centropus rectunguis

Habitat and general habits

International 2001). Many museum specimens registered as C. rectunguis in fact are C. sinensis (e.g., specimens at MZB and RMNH). In peninsular Thailand C. rectunguis are reported at Bala, Khlong Hala (Robson 2000b), and in peninsular Malaysia in 18 localities including Belum Forest Reserve, Taman Negara NP, Tumboh FR, Tekam FR, Dusun Tua, Subang FR, Pasoh FR, Sungai Kinchi, EndauRompin Conservation Area and Rengam FR (Wells 1999, BirdLife International 2001). In Sumatra they are reported at Barat in Padang highlands and southern Barisan Range, Jambi in Ketalo and Tembesi Rivers, and Selatan in Barisan Range (Salvadori 1879, Büttikofer 1887, van Marle and Voous 1988); but most specimen records are old and are unsupported by date, and the upland records are questionable because these coucals are lowland birds (BirdLife International 2001). In Borneo most records are in lowland Sarawak and Brunei, they also occur in Sabah (where they occur in lowlands from sea level to 400 m, as at Danum Valley Conservation Area) and a few lowland sites in Kalimantan (Kutai NP, Sungai Sebangau, Gunung Palung NP) (BirdLife International 2001, Sheldon et al. 2001).They do not occur on nearshore islands. Resident.The coucals are fairly common at low densities in tall lowland forest south of 5o N latitude in the Malay Peninsula (Wells 1999). Their conservation status is considered vulnerable, due to their rapidly disappearing lowland mature forest habitat (BirdLife International 2001).

Ground and understory of lowland closed canopy forest (broad-leafed alluvial forest, peat swamp forest and riverine forest). Occur in lowlands, and sometimes in nearby hills to 600 m. In the Malay Peninsula they are found in mature forest, with a few records from recently logged forest mainly in the lowlands below slopes where logging is less complete. In Sumatra they live in undisturbed and selectively logged primary forest and forest edges. In Sabah (Gunung Mulu) and Kalimantan (Gunung Palung) they occur in lowland and alluvial forest (not swamp or upland forest) (Wells 1976, Laman et al. 1996). Sight and sound records in open habitats in scrub, padi fields and tall grass perhaps apply to Greater Coucal Centropus sinensis (Johns 1989, Holmes 1995, Sheldon et al. 2001). Resident. Solitary and terrestrial, Short-tailed Coucals live on the ground and under-stratum of lowland forest.

Food Few records are known, the coucals eat insects and frogs and take birds from mist nets (Wells 1999, BirdLife International 2001).

Breeding In the Malay Peninsula the coucals call mainly from November to March and also in May and August, there is a nest record in September, and dependent fledglings with an adult were seen in May (Chasen 1939, Medway and Wells 1976, Wells 1999). In Sumatra, a nestling was found in March (van Marle and Voous 1988). The nest is an untidy covered structure of leaves, some still attached to twigs, the nest lined with teased palm-frond pith, built in fronds of a stemless Salacca palm, 2 m above ground in mature forest (Wells 1999). Eggs are white, 37 ⫻ 30 mm (Schönwetter 1964). Clutch size, incubation and nestling periods are unknown. Both parents attend to the nest (Wells 1999).

Bay Coucal Centropus celebensis 221

Bay Coucal Centropus celebensis Quoy and Gaimard, 1830 Centropus celebensis Quoy and Gaimard 1830, Observations zoologiques faites à board ‘de l’Astrolabe , 1, 1, p. 230; Atlas, Oiseaux, pl. 20. (Manado, Celebes) Polytypic. Two subspecies. Centropus celebensis celebensis Quoy and Gaimard, 1830; Centropus celebensis rufescens (Meyer and Wiglesworth, 1896). Other names: Pyrrhocentor celebensis (Quoy and Gaimard 1830).

(34.6 ⫾ 1.9), F 33–41 (38.1 ⫾ 3.0); tarsus, M 36–44 (40.6 ⫾ 2.2), F 38–50 (41.6 ⫾ 3.9), hallux claw, M 16–20 (17.6 ⫾ 1.3), F 17–21 (18.4 ⫾ 1.3), the hallux and hallux claw are short and curved, one M (SMF 28232) had a straight hallux claw (AMNH, BMNH, SMF). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Description

Field characters

ADULT: Sexes alike, head and upper back buffy gray, shaft streaks buff, lower back and rump rufous chestnut, wing and tail rufous chestnut, breast and belly gray to rufous chestnut, under tail coverts rufous chestnut; iris red, bill black, legs and feet black.

Overall length 44–50 cm. Coucal with unmarked light gray to brown plumage, wings and tail dark rufous, bill black.

JUVENILE: Like adult, throat and breast paler, rectrices narrow; iris light gray, lower mandible and tip of upper mandible pale. NESTLING: Undescribed. SOURCES: BMNH, MSNG, MZB, ROM, SMF, UMMZ, ZMUC, ZSM.

Subspecies Centropus celebensis celebensis Quoy and Gaimard, 1830; crown and upper back buffy olive gray, breast dark gray; N Sulawesi; Centropus celebensis rufescens (Meyer and Wiglesworth, 1896); crown buffy brown to rufous, breast dark rufous; C, E, S and SE Sulawesi, Labuan Blanda, Muna and Butung. White and Bruce (1986) considered this coucal to form a superspecies with Rufous Coucal Centropus unirufus, but its song is more like that of Black-faced Coucal C. melanops. C. celebensis, C. unirufus and C. melanops lack the long straight hallux claw of most other coucals.

Measurements Wing, M (n ⫽ 10) 164–184 (173.2 ⫾ 7.1), F (n ⫽ 8) 172–187 (181.6 ⫾ 5.9); tail, M 250–302 (267.1 ⫾ 14.3), F 282–333 (296.0 ⫾ 22.7); bill, M 31–38

Voice Calls include (1) a series of “woop” notes; (2) a series of deep “hoo” notes that accelerate; then the mate joins and the pair duets together, one decreasing the pitch then slowing the phrases and raising the pitch again, the “water bottle” song (Meyer 1879, Holmes and Phillipps 1996). Solitary birds give a repeated “wheeze” (Watling 1983, NSA; Smith 1993a).

Range and status Sulawesi and neighboring islands of Labuan Blanda, Muna and Butung. Resident (Riley 1924, Stresemann 1924, van Bemmel and Voous 1951,White and Bruce 1986).These birds occasionally fly across small treefall gaps in the forest (Coates and Bishop 1997), although larger gaps may be barriers to local dispersal. Sparsely distributed, locally common, they are

222 Gabon Coucal Centropus anselli difficult to observe in dense vegetation.They live in Tangkoko Nature Reserve to 600 m, in Gurung Rurukan to 800 m, in Tentolo-Matinan Mts to 600 m, in Lore Lindu NP to 1000 m, in the Mekonga Mts to 550 m and on the Lompobattang Massif to 1100 m (Coates and Bishop 1997); they are locally common in Dumoga-Bone NP (Rozendaal and Dekker 1989). Rattan, the key habitat factor to the coucals’ survival, is rapidly being stripped from the coucals’ forests.

climbing through lianas and thick foliage (Watling 1983). On Sulawesi they occur in more forested habitat than the Lesser Coucal C. bengalensis (Rozendaal and Dekker 1989).They sometimes forage in small groups (Watling 1983), and they often associate with Yellow-billed Malkoha Rhamphococcyx calyorhynchus and troops of macaques Macaca nigra (Heinrich, in Stresemann 1940,Wardill et al. 1998).

Habitat and general habits

Fruit, including nutmeg (Meyer 1879); large insects, orthoptera, locusts, beetles; and spiders (Stresemann 1940).

Lowland and lower montane rain forest, forest edge and dense regrowth, in lianas and rattan thickets, ranges into secondary lowland and hill forest, open woodland and scrub, also mangrove forest (Rozendaal and Dekker 1989, Coates and Bishop 1997). Mainly in lowland forest, occur from sea level to 1100 m. Arboreal, they live in the undergrowth and lower forest canopy. In behavior this coucal is more like a malkoha than a typically terrestrial coucal,

Food

Breeding A female had an enlarged ovary in February (AMNH 298842). Nest is built of brushwood, flat and open not covered or domed, above ground on trees in forests (Meyer 1879). The eggs, incubation and nestling periods are unknown.

Gabon Coucal Centropus anselli Sharpe, 1874 Centropus anselli Sharpe, 1874, Proceedings of the Zoological Society of London 1874, p. 204, plate 33, fig. 1. (Danger River, Gabon) Monotypic.

Description ADULT: Sexes alike, plumage dark, above head and upper back black glossed purplish, lower back dark brown, wing rufous with darker tips, inner secondaries dark brown, lower back and rump buffy white barred black, upper tail coverts shaggy and long, barred black and buff, tail black with a bronze gloss, T1 narrowly barred buff, the outer tail feathers T2 to T5 unmarked black, hackle shafts black above to ochre on chin to breast, face and side of neck black to well below the eye, chin to belly and under tail coverts rufous buff, under wing coverts rufous buff; iris red to reddish brown, some females with a whitish outer ring, bill black, feet black. Nonbreeding plumage, restricted to the first year, above the head and nape to upper back dark brown with

pale inconspicuous shaft streaks, back unbarred dark brown, face dark brown with irregular buff spots, throat, breast and belly rufous buff with narrow dark brown bars, wing barred as in juvenile. Birds molt into this barred plumage from the juvenile plumage. JUVENILE: Head brownish black with fine pale shaft streaks, upper back black with black shaft streaks, lower back dark brown, wing coverts and wing rufous with dark brown bars, inner secondaries dark brown, rump and upper tail coverts barred buffy white and dark gray, tail black with buff bars, throat and upper breast rufous buff with fine black bars and shafts have buff and black bars, belly unbarred buff, under tail coverts rufous buff with indistinct blackish bars, under wing coverts rufous buff;iris gray brown,bill dark gray above and pale gray below, feet slate to bluish gray. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, UMMZ, ZFMK.

Gabon Coucal Centropus anselli 223

Measurements and weights Wing, M (n ⫽ 11) 178–197 (186.0 ⫾ 7.0), F (n ⫽ 14) 180–209 (194.8 ⫾ 11.8); tail, M 242–302 (266.2 ⫾ 24.5), F 280–380 (302.2 ⫾ 22.7); bill, M 30–37 (33.1 ⫾ 1.9), F 31–38 (34.8 ⫾ 2.2); tarsus, M 40–48 (44.2 ⫾ 2.2), F 45–52 (48.9 ⫾ 2.3); hallux claw, M 17–18 (17.2 ⫾ 0.5), F 16–20 (17.4 ⫾ 1.3) (AMNH, FMNH). Weight, M (n ⫽ 1) 210 (FMNH). Wing formula, P5 ⫽ 4 ⫽ 3 ⫽ 2 ⬎ 1 ⬎ 6 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 46–58 cm.A coucal of central African forest, with crown and sides of head black, back dark brown and underparts rufous buff. Juvenile is very similar to juvenile Black-throated Coucal Centropus leucogaster neumanni, the head more brown and the feathers with fine streaks not spots.

Voice Fast call is a trill, more rapid and more modulated than that of other large African coucals; the note starts at 0.5 kHz then jumps to 0.7 kHz and slides down to 0.4 kHz, repeated in a trill on one pitch at 7–8 notes per sec, slowing to 5 notes per sec near the end (Chappuis 2000). Another call is a slower series of loud, deep notes at 0.6 kHz, the note 0.2–0.3 sec long and repeated 3 in a second, a melancholy bass “ouh ouh ouh ouh . . .”, like the call of the Black-throated Coucal C. leucogaster. Calls are delivered with head lowered and throat puffed out, and pairs call in duet, one bird giving the trill and the other the longer notes (Heinrich 1958, Chappuis 1974, 2000, Irwin 1988, Dowsett and Dowsett-Lemaire 1993, Christy and Clarke 1994).

1997), and an isolated population occurs in northwestern Angola (Dean 2000). Resident. Frequent to common.Territories are a few hectares in area, in dense vegetation where the birds are hard to observe. Population density is c. 3–5 pairs / 100 ha in Gabon, where territories are exclusive to Blueheaded Coucal C. monachus (Brosset and Erard 1986). The range complements that of Blackthroated Coucal C. leucogaster.

Habitat and general habits Primary forest, undergrowth in swampy forest, forest second growth, forest edge along rivers, old cultivation, grassy swamps. Terrestrial, active, they feed mainly on or near ground and they scavenge around camps and villages (Bannerman 1933, Chapin 1939, Irwin 1988, Louette 1986).

Food Omnivorous, taking insects (grasshoppers, katydids, beetles), mollusks, frogs, small snakes, small birds (Bates 1930, Chapin 1939, Mackworth-Praed and Grant 1970, Brosset and Erard 1986, Irwin 1988, AMNH).

Range and status Central Africa from southern Cameroon (Ebolowa, Sangamélima, Abong Mbang, Bitye on River Dja) (Good 1952, Louette 1981a; AMNH, BMNH, CM, FMNH, UMMZ), southwestern Central African Republic (Dzanga-Ndoki NP) (Green and Carroll 1991), Gabon, Congo (north to Nouabele-Ndoki) and central Zaire (Chapin 1939, Snow 1978, Dowsett and Dowsett-Lemaire

Breeding Laying begins with the rains or in the little dry season in the wettest forest region. In NE Gabon, coucals lay in December and February (Brosset and Erard 1986), young are seen in Uele in November (Chapin 1939) and a female had an egg nearly ready to lay in Angola in March (Heinrich 1958; FMNH 224380).The nest and eggs are unknown.

224 Black-throated Coucal Centropus leucogaster

Black-throated Coucal Centropus leucogaster (Leach, 1814) Polophilus leucogaster Leach, 1814, Zoological Miscellany, 1, p. 117, pl. 52. [“New Holland” ⫽ Gold Coast Colony] Other common names: Great Coucal Polytypic. Three subspecies. Centropus leucogaster leucogaster (Leach, 1814); Centropus leucogaster efulensis Sharpe, 1904; Centropus leucogaster neumanni (Alexander 1908).

Centropus leucogaster efulensis Sharpe, 1904; black plumage glossed blue to green, inner secondaries dark olive brown; SW Cameroon; Centropus leucogaster neumanni (Alexander 1908); plumage as in efulensis, smaller; NE Zaire and extreme W Uganda. The small eastern form neumanni has also been considered a distinct species, more closely related to C. anselli than to leucogaster (Louette 1986).

Description ADULT: Sexes alike, above head and upper back black, lower back, rump and upper tail coverts black barred buff, feathers broad, scallop-shaped (not lanceolate as in the other African coucals), wing and wing coverts rufous chestnut, flight feathers darker at tip, tail black glossed blue, often barred buff, underparts black with the lower breast and belly white with the feathers tipped buff, flanks and under tail coverts rufous buff; iris red, bill black, feet black to blue gray. JUVENILE: Head and upper back blackish, feathers spotted with buff base and short buff shaft streaks, shafts black with a subterminal bar of yellowishwhite rump and upper tail coverts black with fine buff bars, wing rufous with dark brown bars, tail black with narrow buff bars, more extensively barred than in adult, throat and upper breast black with small buff spots, lower breast to belly and under tail coverts unbarred whitish to rufous buff; iris gray or brown, bill blackish above, lower mandible horn. NESTLING: Upperparts with dense white hair-like down; tongue has a black U-shaped mark (AMNH, Chapin 1939). SOURCES: AMNH, BMNH, CM, FMNH.

Subspecies Centropus leucogaster leucogaster (Leach, 1814); black plumage glossed violet blue, inner secondaries dark rufous; West Africa from southern Senegal and Guinea-Bissau to Ivory Coast, Ghana and SE Nigeria;

Measurements and weights C. l. leucogaster: Wing, M (n ⫽ 8) 182–207 (190.3 ⫾ 7.4), F (n ⫽ 8) 187–213 (198.4 ⫾ 8.6); tail, M 244–320 (276.5 ⫾ 23.0), F 262–310 (306.0 ⫾ 20.0); bill, M 31–46 (35.1 ⫾ 5.0), F 34–41 (37.1 ⫾ 2.6); tarsus, M 39–46 (45.1 ⫾ 3.6), F 46–54 (51.2), hallux claw, M 15–22 (17.8 ⫾ 2.4), F 16–22 (19.2 ⫾ 2.3) (AMNH, BMNH, FMNH, RMNH); C. l. efulensis: Wing, M (n ⫽ 14) 180–204 (188.4 ⫾ 6.8), F (n ⫽ 6) 197–223 (207.3 ⫾ 10.7), tail, M 254–291 (274.3 ⫾ 10.7), F 240–336 (288.8 ⫾ 33.3) (AMNH, BMNH, FMNH); C. l. neumanni: Wing, M (n ⫽ 3) 174–182 (178.7), F (n ⫽ 1) 176; tail, M 241–268 (252), F 250 (AMNH, BMNH); wing, U (n ⫽ 36) 165–198.5 (179.9) (Louette 1986). Weight, C. l. efulensis: M (n ⫽ 1) 270 (Eisentraut 1963); C. l. neumanni: M (n ⫽ 1) 293; F (n ⫽ 2) 327–346 (336) (BMNH). Wing formula, P6 ⬎ 5 ⫽ 7 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 46–58 cm. The largest African coucal, rufous chestnut above with a black head, mantle and throat and a white belly (underparts rich buff in Gabon Coucal C. anselli).

Voice Series of 8–20 deep “hoo” notes given 4 per sec at 0.4 kHz, held on one pitch, a falling and rising song with the series slowing in tempo and falling in

Black-throated Coucal Centropus leucogaster 225 pitch; calls at night (Bates 1930, Chapin 1939, Chappuis 1974, 2000, Irwin 1988).

Range and status West and Central Africa from southern Senegal (Casamance) and Sierra Leone to Togo, in southern Nigeria and Cameroon, and a disjunct population neumanni in NE Zaire and western Uganda. Resident. Fairly common in forests of lower Casamance, locally common in Liberia, uncommon in Ghana, rare in Togo, and common in the rivers area of southern Nigeria (Chapin 1939, Lamarche 1980, Louette 1986, Irwin 1988, Morel and Morel 1990, Dowsett and Dowsett-Lemaire 1993, Elgood et al. 1994, Rodwell 1996, Barlow et al. 1997, Gatter 1997; BMNH, CM, FMNH). In southern Cameroon it occurs in Korup NP, Kumba, Mt Kupé and Mt Cameroon (Buea, Victoria) and from Yaounde to the lower Sanaga River (Edea), inland at Lolodorf, and south to Kribi and Efulen, while Gabon Coucal C. anselli occurs further south and east (Good 1952, Louette 1981a,Thomas 1991, Bowden 2001); the two closely approach each other with C. leucogaster at Change, 20 miles NW of Ebolowa, and C. anselli at Ebolowa (Good 1952, FMNH). A report of C. leucogaster in the Mts Mandigues in Mali (Lamarche 1980) is questioned (R. J. Dowsett, in Lachenaud 2003); a report in Niger near Niamey (Debout et al. 2000) was rejected (Demey et al. 2001), the habitat there is not forest and the species has never been seen in Parc W (Crisler et al. 2003; C. Jameson, pers. comm.); and a provisional report of C. leucogaster in Gabon (Bannerman 1933) was rejected (Dowsett and Dowsett-Lemaire 1993).

C. l. neumanni occurs in Zaire north of the equator on the Mbomou River, Uele River and the right bank of the Congo River and in western Uganda (Bannerman 1919, Chapin 1939, Snow 1978, Louette 1986, Rossouw and Sacchi 1998).

Habitat and general habits Forest. Dense undergrowth in forest edge, forest remnants, dense gallery forest, secondary forest, thickets, dense grass, tall reeds and sedges; lowland forest zone, dense and swampy areas, along logging roads and Raphia swamps. In dense gallery forest and thick second growth more than in primary forest. In Cameroon they call on farms above Nyasoso bordering the Mt Kupé forests (Bowden 2001). In Uganda they live in Semliki NP (Rossouw and Sacchi 1998). Mainly in lowlands, to 1000 m in Liberia in montane ravines.Terrestrial, they feed on or near the ground (Chapin 1939, Eisentraut 1973, Irwin 1988, Gatter 1997).

Food Insects, including caterpillars, beetles, grasshoppers, mantids, bugs; also spiders, terrestrial snails and frogs (Bannerman 1933, Chapin 1939, Irwin 1988, Gatter 1997).

Breeding In Liberia from November to February with young attended to by parents in March, independent in April; birds call from August to November and are quiet in April; in Ghana a laying bird was caught at Kumasi in August; in Cameroon they breed from June to November, in Zaire (Uele and Ituri Districts) from March to December (Bannerman 1919, Chapin 1939, Bannerman 1951, Serle 1954, Grimes 1987, Irwin 1988, Gatter 1997). The nest is a large ball of leaves and grass, lined with green leaves, placed in a bush up to 30 cm above ground, in forest or in tall grass (Chapin 1939). Eggs are white, elliptical, of unknown size (the record in Mackworth-Praed and Grant 1970 is questioned, M. P. Walters). Clutch is 2. Incubation and nestling periods are unknown.The nest is tended by male (and female?) (Chapin 1939).

226 Senegal Coucal Centropus senegalensis

Senegal Coucal Centropus senegalensis (Linnaeus, 1766) Cuculus senegalensis Linnaeus, 1766, Systema Naturae ed. 12, 1, p. 169. (Senegal) Polytypic. Three subspecies. Centropus senegalensis senegalensis (Linnaeus, 1766); Centropus senegalensis aegyptius (Gmelin, 1788); Centropus senegalensis flecki Reichenow, 1893. Other common names: Rufous-bellied Coucal.

Description ADULT: Sexes alike (female slightly larger), crown and neck black with gloss, feather shafts stiff and shiny black, upper back and wing rusty brown, rump unbarred black, tail blackish with green gloss; underparts white, feather shafts of breast stiff and yellowish, flanks finely barred blackish, under tail coverts buff, under wing coverts light rufous; iris red, bill black, feet gray.A dark rufous colored plumage phase “epomidis” occurs in humid parts of West Africa (Liberia, Ivory Coast, Ghana, Nigeria; Serle 1950a, Elgood 1955, 1973, Louette 1981b, Grimes 1987, Elgood et al. 1994, Gatter 1997, AMNH, BMNH, RMNH) with the crown, face and throat dark brown to black, belly, flanks and under tail coverts all dark rufous. In Nigeria about half the birds are this of plumage form near the coast, with the proportion in rufous plumage decreasing away from the coast, and none further than 200 km from the coast. JUVENILE: Crown dark brown or gray, back rusty brown barred dark brown, wing and wing coverts rufous barred dark brown, primaries unbarred except for dark at tips, secondaries barred dark brown, upper tail coverts long, blackish barred buff, tail blackish with faint buff bars at tip, throat and breast buffy white with prominent shafts straw to brown, belly buffy white without the glossy shafts, flanks buffy barred blackish, under tail coverts dark buff, under wing coverts light rufous, iris olive brown to yellow brown. NESTLING: At hatching, skin black above, dark pink below, with long white hair-like down; bill pale pink, gape pink, tongue red, iris gray, legs gray.

SOURCES: AMNH, BMNH, BWYO, CM, FMNH, MVZ, ROM, UMMZ, USNM, ZMUC.

Subspecies Centropus senegalensis senegalensis (Linnaeus, 1766); above, head black glossed green, back rufous; Senegal and The Gambia to Nigeria, Chad, Somalia, Uganda and W Kenya, south to Zaire; Centropus senegalensis aegyptius (Gmelin, 1788); sootier, crown dull brownish black, back brown not rufous, larger; the Faiyum and Nile of Egypt south to 28°N; Centropus senegalensis flecki (Reichenow, 1893); top of head and neck black with blue gloss, larger than West African birds; southern Africa from Zimbabwe and Botswana to Zambia, southern Zaire, Malawi and SW Tanzania.

Geographic variation Most striking in the rufous-bellied form “epomidis”, once described as a distinct species Centropus epomidis (Bonaparte 1850). This plumage form of C. s. senegalensis is common in humid coastal west Africa, especially in Nigeria where it occurs within 200 km of the coast, and it has been observed to 500 km inland in Ivory Coast (Brunel and Thiollay 1969, Lachenaud 2003). A plumage form with black throat and upper breast and a white belly has been seen in Ivory Coast near Abidjan where both white- and rufous-breasted color phases are common (Lachenaud 2003); this may be the coucal observed near Niamey, Niger, that was identified as another species (Debout et al. 2000). Birds near Lake Chad are sometimes pale, with crown grayish brown and back pale rufous brown, and were once known as a distinct form C. s. tschadensis Reichenow 1915.

Measurements and weights C. s. senegalensis: Wing, M (n ⫽ 9) 150–176.5 (159.7 ⫾ 7.5), F (n ⫽ 10) 155–185 (164.3 ⫾ 10.0); tail, M 179–195 (188.9 ⫾ 11.1), F 172–224 (195.7 ⫾ 16.0); bill, M 28.6–31.8 (29.6 ⫾ 1.7),

Senegal Coucal Centropus senegalensis 227 F 27–33.4 (30.5 ⫾ 2.0); tarsus, M 36.1–40 (37.7 ⫾ 1.7), F 36.5–42 (38.3 ⫾ 1.6); hallux claw, M 16.4–23.4 (19.1 ⫾ 3.3), F 16.4–21.0 (19.6 ⫾ 2.0) (FMNH, UMMZ, ZMUC); form “epomidis”: Wing, M (n ⫽ 3) 161–163 (162), F (n ⫽ 5) 154–166 (160.4 ⫾ 4.5) (BMNH, RMNH); C. s. aegyptius: Wing M (n ⫽ 16) 170–183 (175), F (n ⫽ 13) 177–187 (183) (Irwin 1988); C. s. flecki: Wing, M (n ⫽ 32), 163–176 (169), F (n ⫽ 32) 170–186 (176) (Irwin 1988). Weight, M (n ⫽ 4) 160–178 (169), F (n ⫽ 5) 163–180 (169) (Irwin 1988). Wing formula, P6 ⫽ 5 ⫽ 4 ⬎ 7 ⬎ 3 ⬎ 2 ⫽ 8 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 36–42 cm. Medium-sized coucal, bill more slender than in the larger Coppery-tailed C. cupreicaudus and Blue-headed Coucal C. monachus. Upper tail coverts black in adult, not barred as in adult Burchell’s Coucal C. superciliosus burchellii.

Kordofan) (Lynes 1925, Chapin 1939, Lamarche 1980, Nikolaus 1987, Snow 1978, Irwin 1988). They occur on offshore islands in the Bijagos Archipelago, Guinea-Bissau (Naurois 1969) and formerly on islands in Lake Chad (Bannerman 1951). In Egypt early in the 20th century it was in the Faiyum, Cairo and the Rosetta Nile; now it is in the Nile Delta and Lower Valley (Goodman and Meininger 1989). Common.

Voice

Habitat and general habits

Song a series of deep “coo” notes, the first two coos often shorter than the others, given at a pitch of about 0.5 kHz, the notes vary in length from 0.06 to 0.12 sec, repeated 3–4 per sec, “ouh ouh ouh ouh ouh ouh ouh ouh ouh ouh ouh ouh ouh”, sometimes dropping in pitch near the end of the series; pairs often duet. Other calls are a short single “ouh”, and a sharp series of “guk, guk” notes at 10 per sec given in alarm or excitement. Calls of dark morph “epomidis” are identical to calls of white-bellied coucals in West Africa (Chappuis 1974, 2000, Irwin 1988, Stjernstedt 1993, Christy and Clarke 1994).

Scrub and thickets, coarse grass, edge of reedbeds, dense riverine bush, sugarcane plantations and other farmlands, towns and gardens, palm groves; less closely associated with wet areas than other African coucals (Irwin 1988, Lewis and Pomeroy 1989, Maclean 1993, Christy and Clarke 1994, Cheke and Walsh 1996, Zimmerman et al. 1996). In southwestern Tanzania near Mbeya, and in Zambia and Zimbabwe along the upper Zambezi River from Kazungula to Katombora, this coucal lives along borders of a dry forest belt, in areas where White-browed Coucal C. superciliosus lives in dense low scrub and cane on flats and shore vegetation of small rivers and occasionally appears in low acacias (Ripley and Heinrich 1966b; RBP). In southern Africa they live in disturbed habitats, including sugarcane fields and gardens (Rowan 1983).They feed mainly on the ground, taking insects from dung of cattle, buffalo and elephants and more dispersed sources. They move in a slow stalking walk then hop and run, and they forage on escaping insects along the edge of grass fire. Flight is low and clumsy as the bird takes a few wingbeats and a glide, then

Range and status Africa south of the Sahara and to Egypt along the Nile. In West Africa they occur from Senegal to Nigeria and Cameroon, in the east from S Sudan and Ethiopia to Lake Victoria and through Central Africa from Uganda to Zaire and N Angola, and in southern central Africa from Zambia and Kenya south to Zimbabwe, the Okavango, Tanzania and S Mozambique. Resident in most of its range, locally migratory in the arid regions (Mali and west

228 Blue-headed Coucal Centropus monachus a crash landing. They sunbathe, perched with back to the sun, back feathers raised, wings drooped and tail spread wide (Hamling 1937, Rowan 1983).

Food Insects, mainly grasshoppers, caterpillars, termites, beetles and bugs; also frogs, small rodents, small snakes, lizards, birds, bird eggs and nestlings; snails (Chapin 1939, Cramp 1985, Irwin 1988).They take distasteful bush locusts Phymantous viridipes, a brightly colored insect with a waxy appearance that shows off its colors and emits a pungent foul liquid, an insect avoided even by ants (Goodwin 2001). Take live birds including Red-billed Queleas Quelea quelea (Ewbank 1985) and scavenge dead fish (Steyn 1970).

Displays and breeding behavior Territorial, the pair duets and advertises its area of c. 6 ha.

Breeding and life cycle During the rains while grass is pliant enough to build the nest and high enough to provide support and concealment. In Egypt they breed from March to August (Goodman and Meininger 1989), in northern Senegal from May to October (Morel and Morel 1990), in The Gambia from July to November (Gore 1990), in Liberia March to April (Gatter 1997), in Mali on the Bani River in August (Bannerman 1951), in northern Ghana July to August, in

southern Ghana at Accra and Cape Coast from March to June (Greig-Smith 1977, Grimes 1987), in Togo they care for fledged young in July (Cheke and Walsh 1996), in Nigeria they breed from March to August (Elgood et al. 1994), in Uele from September to November (Chapin 1939), in Ethiopia from April to August (Urban and Brown 1971), in Kenya from March to May (Brown and Britton 1980), in Malawi from November to May (Benson and Benson 1977), and in Zimbabwe and Botswana from October to March (Irwin 1981, Skinner 1996, Vernon et al. 1997). Birds call in duet nearly all year. The nest is a ball of coarse dry grass, lined with green leaves, the base of the nest more substantial than the dome, to 4 m above ground in bush.A nest I found in The Gambia in 1999 was open, without a cover built by the bird: the nest was built just below the crown of a citrus tree that provided a natural cover of dense spiny branches and leaves. Eggs are white, 34 ⫻ 26 mm, laid in a clutch of 2–5 (Granvik 1923, Bannerman 1933, 1951, Chapin 1939, Irwin 1988, Tarboton 2001). The incubation period is 17–19 days. Nestlings hatch asynchronously, and both parents give parental care. The nestlings are 24 g at day 3, half fledging weight by 8 days, and their wing feathers and ventral contour feathers burst their sheaths by 12 days. The young fledge in 18–20 days at a weight of 125–145 g, well before they can fly; their legs are well developed and they move about in dense vegetation (Steyn 1972).

Blue-headed Coucal Centropus monachus Rüppell, 1837 Centropus monachus Rüppell,1837,Neue Wirbelthiere zu der Fauna von Abyssinien gehorig, entdeckt und beschrieben, Vögel, p. 57, pl. 21, f. 2. (Kulla, northern Ethiopia) Polytypic. Three subspecies. Centropus monachus monachus Rüppell, 1837; Centropus monachus fischeri, Reichenow 1887; Centropus monachus occidentalis Neumann, 1908.

Description ADULT: Sexes alike, above, forehead and crown to upper back black with blue gloss, shaft of the hackles black, lower back and wing reddish brown, rump unbarred black, upper tail coverts either

barred or unbarred, tail black with greenish or bronze gloss, throat to belly and under tail coverts whitish to pale buff, darker buff on flanks; iris dark red, bill black, feet black. JUVENILE: Crown and neck dull blackish with narrow buff shaft streaks, back barred rufous and dark brown, wing rufous barred blackish, rump and upper tail coverts blackish with narrow buff bars, tail black narrowly barred buff, the central rectrices T1 either unbarred or barred near the tip, chin and throat buffy white, breast darker pinkish buff with small black spots and buff shaft streaks, belly buffy white.

Blue-headed Coucal Centropus monachus 229 NESTLING: At hatching, skin is black with long white hair-like down. SOURCES: AMNH, BMNH, CM, FMNH, MAC, KMMA, ROM, ZFMK, ZMB.

Subspecies Centropus monachus occidentalis Neumann, 1908; back and wing rufous brown, larger than C. m. fischeri; West Africa from Liberia and Ivory Coast to Cameroon, Gabon, N Angola and Zaire; Centropus monachus fischeri Reichenow, 1887; adult back and wing dark chestnut, juvenile darker than other populations; Upper Nile basin in S Sudan, Uganda,W Kenya (Lake Victoria basin) and NW Tanzania, Burundi and Rwanda; Centropus monachus monachus Rüppell, 1837; back and wing rufous, brighter than other forms; Eritrea and Ethiopia to C Kenya.

Geographic variation Additional subspecies have been described for Centropus monachus and the plumage variation with age is peculiar to certain forms. The form heuglini Neumann, 1911, with the head dull violet-blue black, in the large permanent swamps of the upper Nile in Sudan, is regarded as synonymous with C. m. fischeri and may be an immature C. m. fischeri (Friedmann 1930). Specimens in ZFMK in the series from southern Sudan taken by A. Koenig and his associates from 1910 and 1913 (including part of the type series of heuglini (van den Elzen and Rheinwald 1984) suggest a distinct subadult plumage; one specimen (CI.3.b´.␥) has a dull black crown (without pale shaft streaks), dark back, one long barred juvenile rectrix and the other (shorter) rectrices dull black, and the rufous wing feathers are incompletely barred, so the bird is not a juvenile and appears to be subadult. The holotype as recognized by van den Elzen and Rheinwald, and described by Neumann and illustrated in color (Koenig 1911) is in Stuttgart, SMNS 5662. Others in the Koenig series are the same in size and in adult plumage, and have a bright purple-blue gloss on the crown and nape.These birds are smaller than C. m. occidentalis in Cameroon, Gabon and Sudan.A form C. m. verheyeni Louette 1986, from Mabwe,

Upemba, in S Zaire was described as smaller, but birds in this locality overlap in size with populations of C. m. fischeri, with which this form may be associated. More work is needed on these coucals to determine their plumage sequence and their relationships to each other.

Measurements and weights C. m. occidentalis, Cameroon, Gabon and Zaire: Wing, M (n ⫽ 8) 180–191 (186.8 ⫾ 4.5), F (n ⫽ 6) 176–196 (185.5 ⫾ 7.1); tail, M 200–228 (216.0 ⫾ 8.4), F 210–224 (217.2 ⫾ 6.2); bill, M 30–34 (31.8 ⫾ 1.2), F 32–35 (33.0 ⫾ 1.1); tarsus, M 43–47 (45.3 ⫾ 2.0), F 43–48 (44.5 ⫾ 1.9), hallux claw, M 18–22 (20.5 ⫾ 2.0), F 21–28 (23.2 ⫾ 2.6) (AMNH); C. m. fischeri (“heuglini” ), southern Sudan: Wing, M (n ⫽ 10) 159–171 (166.5 ⫾ 4.7), F (n ⫽ 5) 167–177 (170.8 ⫾ 4.1); tail, M 181–223 (197.8 ⫾ 11.5), F 188–230 (290.8 ⫾ 15.6) (ZFMK); C. m. monachus, Ethiopia and C Kenya, wing, M (n⫽ 3) 188–220 (201.7), F (n⫽ 14) 182–226 (198.6⫾ 12.3); tail, M 218–237 (225.7), F 208–253 (229.3⫾ 13.3) (AMNH, FMNH). Weight, M (n ⫽ 5) 163–177 (171), F (n ⫽ 3) 206–283 (237) (BMNH). Wing formula, P6 ⫽ 7 ⱖ 5 ⬎ 4 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 45–52 cm. Coucal with large bill, crown and upper back black, back and wings rufous-brown, rump dark, unbarred in adult (barred in the larger Coppery-tailed Coucal C. cupreicaudus), tail long with deep bronze sheen, below white in the adult, bill large and black. Differs from White-browed Coucal C. s. superciliosus in unstreaked head. The immature has a rich buff breast and is more broadly barred on the wing than in the Senegal Coucal C. senegalensis; immature C. cupreicaudus is not barred on the wing.

Voice Deep and resonant, a slow “hoo, hoo, hoo-wu-wuwu-wu-wu hoo hoo, hu” series, the first two notes short and given with a pause before the rest of the series, the first at 0.48 kHz and the later notes

230 Blue-headed Coucal Centropus monachus slightly lower (0.43 kHz) in pitch, the later notes 0.10–0.12 sec long given at a constant rate of 4 per sec. Pairs often duet. Other calls include a monkeylike bark and a raucous cackle (Bates 1930, Chappuis 1974, 2000, Irwin 1988, Christy and Clarke 1994, Zimmerman et al. 1996).

Range and status Africa from Liberia and Ivory Coast to Cameroon, Gabon, southern Sudan, Ethiopia, western Kenya, Uganda, and east and southern Zaire (Friedmann 1930, 1966, Chapin 1939, van Someren 1949, Ripley and Heinrich 1966a, Traylor 1960, Urban and Brown 1971, Snow 1978, Louette 1986, Nikolaus 1987, Irwin 1988, Morel and Morel 1990, Gatter 1997). Resident. Common in village areas in NE Gabon (Christy and Clarke 1994) and E Zaire (Kizungu 2000) and in savannas of N and E Zaire, and in western Uganda around Lake Victoria and Lake Kyoga (Lewis and Pomeroy 1989, Rossouw and Sacchi 1998).

Habitat and general habits Swamps, high grass, marshes, papyrus and river banks, forest edge and second growth, mesic savannas near water or forest, dense cover, sugar cane fields, old manioc fields, edges of villages and along trails (Bannerman 1933, Chapin 1939, Thiollay

1985, Brosset and Erard 1986, Grimes 1987, Zimmerman et al. 1996). Occur in lowlands to 1000 m in lower montane regions in western Cameroon (Eisentraut 1973, Bowden 2001), to 2000 m and 2500 m in Kivu volcanoes (Chapin 1939) and to 2000 m in Kenya where they are within the 1000⫹ mm rainfall region (Lewis and Pomeroy 1989), in wetter habitats than C. senegalensis. Creep about on ground and through dense vegetation.

Food Generalist carnivore; takes insects, mainly grasshoppers and beetles, also snails, frogs, lizards and snakes, small birds (scrub warbler Bradypterus and reed warbler Acrocephalus, swallowed whole), nestling birds, bird eggs, mice and rats (Bates 1930, Chapin 1939, Verheyen 1953, Irwin 1988). One coucal killed a four-foot long cobra by attacking its head (Bates 1909). In captivity, they take slugs (Royston 1981).

Breeding In Togo they breed in June and July (Cheke and Walsh 1996), in Nigeria from March to June (Elgood et al. 1994), in Cameroon in January and from April to September (Bates 1909, Serle 1950b, Mackworth-Praed and Grant 1970, Irwin 1988), in Gabon from August to March (Christy and Clarke 1994), around Lake Victoria in September and from February to June (Brown and Britton 1980), in Zaire in Uele District from May to November (Chapin 1939,AMNH) and in Irangi in December (Kizungu 2000). The nest is an oval mass of dry grass and sedges (or sticks and dry leaves), lined with green leaves, with a side entrance, concealed in bushes or tall grass or a dense tree, placed about 0.2 to 3 m above ground. The eggs are white, with a light gloss, 34.5 ⫻ 27 mm, clutch 3–4 (Bates 1909, 1930, Bannerman 1933, Chapin 1939, Brosset and Erard 1986, Kizungu 2000). The male attends the nest when it has eggs (AMNH 159145, 627794). The incubation and nestling periods are unknown.

Coppery-tailed Coucal Centropus cupreicaudus 231

Coppery-tailed Coucal Centropus cupreicaudus Reichenow, 1896 Centropus cupreicaudus Reichenow, 1896, Ornithologische Monatsberichte, 4, p. 53. (Okawangoland and Angola ⫽ S Angola] Polytypic. Two subspecies, Centropus cupreicaudus cupreicaudus Reichenow 1896; Centropus cupreicaudus songweensis Benson 1948.

Description ADULT: Sexes alike, above the forehead to upper back black with violet gloss, back dark brown, wing and wing coverts reddish brown, rump and upper tail coverts black, often finely barred, tail blackish brown with coppery gloss; underparts creamy white; iris red, bill black, deep in shape, feet black. JUVENILE: Head dull black with black shaft streaks, back dark brown, wing reddish brown barred dark brown, rump and upper tail coverts blackish, tail blackish with fine buff tip, chin to belly whitish, breast has buff shaft streaks and small blackish spots. NESTLING: At hatching the skin is black, chin and abdomen tan, egg tooth white; long hair-like white down on head and spinal tract, shorter gray hairlike down on the wing and legs. SOURCES: AMNH, BWYO, BMNH, FMNH, UMMZ.

Subspecies Centropus cupreicaudus cupreicaudus Reichenow 1896; larger and paler; Angola to Zambia and Zimbabwe; Centropus cupreicaudus songweensis Benson 1948; smaller and darker; southern Tanzania and northern Malawi.

(41.1 ⫾ 1.5), F 41.4–43.0 (42.4 ⫾ 0.7); tarsus, M 47.0–57.6 (51.2 ⫾ 3.8), F 44.0–53.2 (49.1 ⫾ 2.7); hallux claw, M 21.8–30.4 (27.5 ⫾ 3.0), F 26.2–30.0 (27.2 ⫾ 2.3) (FMNH); C. c. songweensis: Wing, MF 195–203 (Irwin 1988). Weight, M (n ⫽ 5) 250–293 (272), F (n ⫽ 5) 245–342 (299) (Irwin 1988); M (n ⫽ 7) 203–228 (216.0), F (n ⫽ 6) 218–233 (223.0) (FMNH). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 8 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 42–50 cm. Large coucal with large bill, crown and upper back black, back dark rufousbrown and wings rufous, rump dark brown usually with indistinct buff bars (unbarred in the smaller Blue-headed Coucal C. monachus), tail long and deep copper; underparts white. Head and neck with violet (not blue) gloss. Juvenile differs from juvenile C. monachus in the lack of barring above and in whitish not buff breast. Wing with dark trailing edge (not dark in C. senegalensis). Where it occurs with Burchell’s Coucal C. s. burchellii, Coppery-tailed Coucal has a larger size and deeper voice; where it occurs with the smaller Whitebrowed Coucal C. s. superciliosus, it has an unstreaked head.

Voice Calls a series of low “coo” notes, at 0.4 kHz either on one pitch or descending in pitch, the note 0.12–0.20 sec long given 3–4 per sec, similar to Blue-headed Coucal C. monachus; contact notes a low “cou . . . cou . . .”. The pair often duets (Chappuis 1974, 2000, Stjernstedt 1993).

Measurements and weights C. c. cupreicaudus; Zambia and Botswana: Wing, M (n ⫽ 7) 202–224 (214.0 ⫾ 8.9), F (n ⫽ 7) 216–236 (224.4 ⫾ 7.6); tail, M 220–263 (245.1 ⫾ 17.9), F 243–273 (254.2 ⫾ 10.7); bill, M 38.8–42.5

Range and status South-Central Africa from Angola, southern Zaire, Zambia, Okavango, Boteti, Linyati and Chobe rivers in Botswana and E Caprivi, upper Zambezi

232 Coppery-tailed Coucal Centropus cupreicaudus

Food Large insects, mainly grasshoppers; also snails, crabs, fish, frogs (including painted reed frog Hyperolius viridiflavis and grass frog Ptychadena sp.), snakes, lizards, small birds (Ploceus weavers, Coturnix quail taken in scavenging), small rodents; also green grass and waterweed. Scavenge for dead fish, and tear open bird nests to take the nestlings (Vincent 1946, Rowan 1983, Irwin 1988).

Breeding River east to Victoria Falls in Zimbabwe, and there are a few records in Malawi and Tanzania. Sparsely distributed but locally common, as in marshes along streams and rivers near upper Zambezi. This species and Blue-headed Coucal C. monachus are allopatric through most of their composite range (Traylor 1960, Snow 1978, Irwin 1988, Louette 1986). Resident. Local movements occur with burning or heavy grazing of wetlands. Birds disappear and reappear when water and wetlands return, perhaps persisting in riparian thicket refuges (Vernon et al. 1997).

Habitat and general habits Wetlands in marshes, swamps, papyrus Cyperus papyrus, reedbeds Phragmites, river bank vegetation, thickets, long rank grass, floodplains and adjacent bush, dambos in miombo woodland. Occurs in lowlands, to 1250 m in Angola. They walk on the ground, feeding on insects, the birds alone or in pairs. Breeding adults search mainly near the nest in long-awned water grass Echinochloa stagnina on floodplain; they also forage up to 1 km from the nest, and occasionally they enter woodland adjacent to the floodplain (Hustler 1997a). A few hundred meters from water and marsh, this species and White-browed Coucal C. superciliosus are replaced by Senegal Coucal C. senegalensis.

In Zaire December to February (Vincent 1946, Irwin 1988), in Zambia during the rains in September and November to February (Benson et al. 1971), in Zimbabwe during the rains of January to April (Hustler et al. 1996, Hustler 1997a). Nest is a coarsely built sphere of fresh grass, coarse straws or twigs, lined with leaves, an entrance on the side, sometimes with a runway of flattened grass to the entrance, built low in reedbeds within 0.5 m of water level, on a square meters few of high ground overgrown with reeds or in a tuft of grass above water. Eggs are white, with little gloss, 38 ⫻ 28 mm, clutch 2–4, sometimes laid before the nest is complete. Incubation period is unknown; it probably begins when the first egg is laid, as the young hatch asynchronously and are of different sizes in a brood (Keith and Vernon 1969). Nestlings are fed locusts and frogs. Both parents provide care to the brood, sometimes regurgitating water for the young to drink. The nestling period is at least 17 days. The young leave the nest when well feathered but before they can fly, and after fledging they are cared for by the adults for at least 40 days. The young beg by wing fluttering and incessant “chuck” calls. The young develop the blue sheen on the nape plumage when they still have a short tail, dark eye and broad barring on some flight feathers. Nest predators include monitor lizards Varanus sp. and water mongoose Atilax paludinosus (Benson and Pitman 1964, Irwin 1988, Maclean 1993, Hustler 1997a).

White-browed Coucal Centropus superciliosus 233

White-browed Coucal Centropus superciliosus Hemprich and Ehrenberg, 1833 Centropus superciliosus Hemprich and Ehrenberg, 1833, Symbolae Physicae, 2, Aves, vol. 1, un-numbered p. 35, footnote 3. (Arabia and Ethiopia) Other common names: Burchell’s Coucal, Lark-heeled Cuckoo. Polytypic. Five subspecies. Centropus superciliosus superciliosus Hemprich and Ehrenberg, 1833; Centropus superciliosus burchellii Swainson, 1838; Centropus superciliosus fasciipygialis Reichenow, 1898; Centropus superciliosus loandae Grant, 1915; Centropus superciliosus sokotrae Grant, 1915.

Description ADULT: Sexes alike, upperparts blackish brown, head with black shaft streaks on hackles, upper back blackish streaked whitish, back rufous brown, rump and upper tail coverts finely barred black and buff, wing rufous chestnut, long and broad tail black glossed green with a thin white band on tip of each tail feather in fresh plumage, face with broad white superciliary streak (lacking in terminal adult plumage of C. s. fasciipygialis and C. s. burchellii); underparts whitish, breast with yellow strawcolored hackle shaft streaks, flanks and under tail coverts buff finely barred black, under wing coverts rufous buff; iris red, bill black, feet bluish gray to black. JUVENILE: Head and neck streaked buff and brown, back barred rufous and blackish, rump and upper tail coverts blackish barred buff, wing rufous chestnut with blackish bars on the tips, tail black with buff bars, face brown with yellow shaft streaks and an indistinct whitish superciliary streak; underparts whitish with buff yellow shaft streaks on throat and breast, belly whitish, flanks and legs buffy gray barred gray; iris brown to red (by the age when the tail is grown). NESTLING: At hatching skin black with long white hair-like down, like a spiny hedgehog. At fledging the young retains the hair-like down on the forehead, crown and neck, and the tail is very short.

SOURCES: AMNH, BMNH, BWYO, CM, FMNH, MSNG, MVZ, ROM, UMMZ, USNM, UWBM, ZFMK, ZMUC.

History and subspecies Centropus superciliosus superciliosus Hemprich and Ehrenberg, 1833; paler on crown and back, smaller than other geographic forms, flanks and under tail coverts barred, only center of belly unmarked white; juvenile more buffy (paler); Ethiopia, W Somalia, E Sudan, Kenya, NE Uganda and NE Tanzania; Centropus superciliosus sokotrae Grant, 1915; paler and grayer, pale shaft to dorsal hackles, not all individuals are distinct from Ethiopian superciliosus; southern Arabian Peninsula, E Somalia and Socotra I; Centropus superciliosus loandae Grant, 1915; larger and plumage darker on crown and back than superciliosus, juvenile crown black with rufous streaks; East Africa from Uganda, SW Kenya, Tanzania, SE Zaire, N Malawi, Zambia and Angola to N Zimbabwe and Botswana; Centropus superciliosus fasciipygialis Reichenow, 1898; crown, head and nape black, no superciliary streak, plumage similar to C. s. burchellii but lacks an intermediate age plumage and size smaller; eastern Zimbabwe, Tanzania, Mafia Island and Mozambique south to Beira; Centropus superciliosus burchellii Swainson, 1838; crown, head and nape black, no superciliary streak, black shaft to dorsal hackles, no apparent streaks on back, underparts white with no melanin in breast hackles; also an intermediate age plumage with pale buff superciliary streak, crown brown, neck blackish with straw-colored streaks; juvenile much more rufous than in other geographic forms, crown blackish with dark rufous spots and shafts, cheeks black, no light line over the eye, throat to breast, flanks and under tail coverts light rufous (not buff ) with indistinct blackish bars on the flanks, center of belly whitish; southern Malawi, southern Zambia, southern Zimbabwe, eastern Botswana and South Africa.

234 White-browed Coucal Centropus superciliosus The geographic complex of populations is considered a single biological species and the southern birds are White-browed Coucals without a white brow. Although the unbrowed forms might be recognized as a species C. burchellii (as in Clancey 1989), birds in intermediate plumage with a partial brow, apparent hybrids between white-browed and unbrowed forms, occur in the regions where these forms approach each other in Tanzania, Zambia, Mozambique and northern Malawi.The difference in plumage suggests that adults of northern and western regions have a juvenile-like plumage with white superciliary streak and brown crown and hind neck, while adults in the south have a glossy black head and neck. The terminal adult plumage of southern birds C. s. burchellii appears to be an extra plumage that is grown after the streaked adult plumage; birds breed in both streaked and unstreaked plumage (Lawson 1962, Snow 1978). In NW Zimbabwe along the upper Zambezi River near Kazungula, in 1997 and 1998 I saw mixed pairs mate and breed, one bird with a partial white eye streak and its mate without a streak. Birds in eastern Tanzania, C. s. fasciipygialis (described as a species fasciipygialis by Reichenow 1898, and considered a form of Centropus senegalensis by Friedmann and Loveridge 1937) have a blackish head without a superciliary streak (as in burchellii), and are small (as in loandae); their plumage varies along the coast with a trace of an eye streak in some birds at Mikindani, with different amounts of buff on the belly, and with a dark submarginal edge to the breast feathers giving a streaked breast (as in nominate superciliosus) that increases in the intensity of striping in the Saga Hills and Same District (ZMUC). Nominate superciliosus and C. s. loandae in East Africa and Zaire differ only slightly in plumage (Chapin 1939, Louette 1986). Because birds vary in plumage within a local population and intergrade with regionally neighboring forms, interbreed where they come together, and have the same behavior, they are considered conspecific.

Measurements and weights C. s. superciliosus; Kenya: Wing, M (n ⫽ 6) 135–152 (145.8 ⫾ 7.5), F (n ⫽ 6) 148–158 (154.2 ⫾ 4.2); tail, M 176–200 (189.3 ⫾ 8.5), F 185–207

(197.3 ⫾ 7.5); bill, M 27.7–29.3 (28.8 ⫾ 0.6), F 28.6–33.3 (31.3 ⫾ 1.8); tarsus, M 28.9–38 (33.3 ⫾ 3.7), F 33.7–38.1 (36.0 ⫾ 1.5); hallux claw, M 15–18.5 (16.9 ⫾ 1.8), F 14.8–21 (18.4 ⫾ 2.0) (FMNH); C. s. sokotrae: Wing, M (n ⫽ 6) 149–160 (155.2 ⫾ 4.0), F (n ⫽ 2) 165–168 (166.5) (MSNG, USNM); C. s. loandae: Wing, M (n ⫽ 6) 157–168 (161.5 ⫾ 3.7), F (n ⫽ 10) 160–172 (167.4 ⫾ 8.2) (FMNH); C. s. fasciipygialis: Wing, M (n ⫽ 6) 149–165 (156.4 ⫾ 6.0), F (n ⫽ 4) 156–166 (161.5 ⫾ 4.4) (ZMUC, UMMZ); C. s. burchellii: Wing, M (n ⫽ 11) 158–179 (168.1 ⫾ 6.3), F (n ⫽ 9) 166–182 (172.6 ⫾ 5.7) (AMNH, BMNH, FMNH, USNM). Weight, C. s. superciliosus: M (n ⫽ 6) 110–140 (124), F (n ⫽ 7) 125–170 (136) (Irwin 1988); C. s. loandae: M (n ⫽ 14) 142–213 (160), F (n ⫽ 10) 153–212 (180) (Irwin 1988); C. s. burchellii: F (n ⫽ 1) 159 (UWBM). Wing formula, P7 ⱖ 6 ⱖ 5 ⬎ 4 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 36–42 cm. Coucal with rufous back and wings and a long tail, whitish below, in marsh and on the ground.White-browed coucals in most of their range have a whitish superciliary streak in adult plumage, in contrast to all other coucals. Burchell’s Coucal C. s. burchellii are similar to Senegal Coucal C. senegalensis except they have a finely barred rump (black rump in C. senegalensis).

Voice First song type is a rapid series of bubbling “coo” notes repeated about 8 per sec, each note at 0.4–0.5 kHz, the song lasting two or three seconds. The first note is longer than the other notes in the series. The series often descends in pitch and increases in tempo, like water gurgling from a narrow-necked bottle. Another bubbling song type is a slower more deliberate series of notes beginning at 0.5 kHz and dropping to 0.4 kHz then rising to 0.5 kHz, the first notes at 6 per sec, the low middle at 9 per sec and the end at 4 per sec, slowing then rising at the end.A single bubble call is given to the

White-browed Coucal Centropus superciliosus 235 mate. Members of a pair call in a duet; one begins as the other completes its call and they often overlap. Excitement or threat call is a rapid chatter of sharp notes given 6 or 7 per sec. Alarm and excitement calls are a hiss and a single low “chuck”. The songs and calls are the same in C. s. burchellii in South Africa, C. s. superciliosus in Kenya and C. s. loandae in Zambia (North 1958, Stannard 1966, Chappuis 1974, 2000, Gillard 1987, Gibbon 1991, Stjernstedt 1993, Christy and Clarke 1994, Skead 1995, Stevenson and Fanshawe 2002, RBP).

Range and status Africa and the Arabian Peninsula (Ogilvie-Grant and Forbes 1903, Ripley and Bond 1966, Cornwallis and Porter 1982, Porter et al. 1987, Kirwan et al. 1996, Stevenson and Fanshawe 2002). They occur in the southern Arabian Peninsula in Yemen south of 17°N from the eastern Tihamah and Jabal Bura east to the western Hadhramaut around 48°E (Meinertzhagen 1954, Brooks et al. 1987), on Socotra Island (Ripley and Bond 1966), and widespread in Africa from Eritrea and Ethiopia, Sudan along the Nile, Zaire, and East Africa to Angola and South Africa. Resident: Locally they sometimes appear with the rains (Lewis and Pomeroy 1989). Common in much of the range (Chapin 1939, van Someren 1956).

Habitat and general habits Riverside, dense bush, scrub, acacia savanna and thickets, reeds and papyrus, moist vegetation, tall rank grass (Pitman 1928), marshes, shrubby once-

cultivated old fields, along watercourses and lush scrub and cultivation in arid country; mostly lowland, also montane areas, in foothills to 1000 m in Yemen ( Jennings 1981, Brooks et al. 1987) and in bracken briar as high as 2800 m in East Africa (van Someren 1956), and cultivated irrigated habitats in arid country. In regions where Senegal Coucal C. senegalensis also occur, C. superciliosus live in more marshy habitat. Solitary, retiring and skulking in behavior, often hidden, seen when they perch on a bush and spread wings and tail to dry and to sun. Flight is weak and horizontal, flapping then gliding, then crashing into the vegetation.

Food Insects, mainly grasshoppers, crickets and locusts, beetles (Curculionidae), also ants and other hymenoptera; spiders, land and garden snails, crabs, scorpions, lizards, snakes, frogs, mice, and small birds including nestlings and eggs (Ripley and Bond 1966, Johnson 1968, Maclean 1983, Rowan 1983, Frere 1984, Irwin 1988, Skead 1995). Small prey are swallowed whole; larger prey are broken into pieces or pulped by pecking, and snails are beaten against a stone. They also take fruit including loquats and buffalo thorn Ziziphus mucronata (Rowan 1983).

Displays and breeding behavior Pairs are monogamous and territorial. Male feeds the female a large insect at copulation (Brooke et al. 1990).

Breeding and life cycle Breeding season, during the rains. In Ethiopia they breed from March to June (mainly in April and May) (Friedmann 1930, Urban and Brown 1980), in southern Somalia fledglings appear in October (Ash and Miskell 1998), in Uganda in S Ankole they breed in October (Pitman 1928), in Kenya they breed in all months, mainly in the wet seasons as in Ngong where they breed from April to July and August (van Someren 1956, Brown and Britton 1980, Zimmerman et al. 1996), in Zanzibar and Pemba there are two distinct breeding seasons, the April to July and the November to January monsoons (Pakenham

236 Javan Coucal Centropus nigrorufus 1979), in Tanzania they breed from January to March (Schuster 1926), in Zambia from December to February (Benson et al. 1971), in Malawi from October to March (Benson and Benson 1977), in South Africa from September to February (Maclean 1993, Skead 1995), and in the SW Cape from August to January (Rowan 1983, Skinner 1996,Vernon et al. 1997). The nest is a large bulky untidy dome with a side entrance, built of grass and twigs, sometimes grass alone, without a distinct cup, the inner surface lined with leaves. It is built low in reeds or bush, from ground level to 10 m high. Nests built of grass are covered with the grass that is twisted around the top, or are built in a bower with several live twigs in the supporting shrub bent and woven over the nest, or in a bush, hedge or thorny tree (Pitman 1928,Tarboton 2001). Eggs are dull white, becoming stained yellow after a few days in the nest, 34.6 ⫻ 22.8 mm (C. s. superciliosus, Sudan, Egypt), 34.5 ⫻ 26 mm (C. s. burchellii, (Irwin 1988)), 32.9 ⫻ 24.7 mm (C. s. loandae, Uganda, (Pitman 1928)). Clutch size is 3–5(6) (Irwin 1988), in

Uganda 2–3 (once 4) (Pitman 1928), in Tanzania 3–4 (Schuster 1926), in Natal 2–5 (4 most common) (Dean 1971), in Eastern Cape 4 (Skead 1995). Eggs are laid at intervals of 2–3 days (Skead 1995). Incubation is mainly by male, beginning with the first egg or later.The incubation period is 14–15 days. Nestlings hatch asynchronously, a day apart. Both sexes, mainly the male, feed the young. The nest accumulates layers of feces (the earliest feces are in capsule form), cast quill-sheaths of growing nestlings, and remains of eggshells on the bottom layer; fresh green leaves are layered over this detritus (Skead 1995). Nestlings hiss and emit a foul black cloacal liquid when disturbed.They fledge at 18–20 days (or as early as 14 days if disturbed), and the older young stay near the nest until all are ready to leave the nest. Fledgling coucals are barely able to fly. They creep about quietly under cover until parents arrive with food (van Someren 1956, Johnson 1968, Irwin 1988, Skead 1995).Young are said to be flown to safety, the parent holding a chick between the parent’s thighs, saving it from an advancing fire (Bannerman 1933).

Javan Coucal Centropus nigrorufus (Cuvier, 1817) Cuculus nigrorufus Cuvier, 1817, Regnè Animale 1, 1817 (1816), p. 426. ( Java) Other common names, Sunda Coucal. Monotypic.

Description ADULT: Sexes alike, female slightly larger, upperparts glossy black with stiff hackles, upper back with a purplish gloss, wing coverts rufous with black ends on inner vane, inner wing coverts black, wing dark rufous with blackish inner vanes and black tips, inner secondaries black, tail long and black, below black glossed purplish; iris red, bill black, feet black. JUVENILE: Plumage like adult. NESTLING: Undescribed. SOURCES: MZB, RMNH, USNM.

Measurements Wing, M (n ⫽ 9) 195–218 (one “M” 238 not included) (207.4 ⫾ 6.4), F (n ⫽ 4) 210–227 (220.8 ); tail, M 182–242 (221.1 ⫾ 20.7), F 242–262 (250.3 ⫾ 10.4); bill, M 37–41 (39.0 ⫾ 1.5), F 40–41 (40.7 ⫾ 0.6); tarsus, M 57–60 (58.0 ⫾ 1.7), F 57–70 (62.3 ⫾ 6.8); hallux claw, M 26–42 (29.5 ⫾ 6.2), F 25–40 (32.0 ⫾ 8.2) (RMNH, USNM). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 46 cm. Coucal with black upperparts glossed purple and stiff hackles, wings rufous with black edges and black areas on the coverts, and a long tail. Differs from Greater Coucal C. sinensis in black back (rufous in C. sinensis), purplish gloss, darker rufous of wings and black inner vanes of the wing flight feathers and coverts.

Javan Coucal Centropus nigrorufus 237

Voice Not recorded or described, said to be similar to C. sinensis (BirdLife International 2001).

Range and status Java. Resident in flat coastal swamps and wetlands: Muara Angke and Muara Gembong near Jakarta, Cangkring and Muara Cimanuk near Indramayu, and Ujung Pangkah and Sidoarjo on the Brantas river delta near Surabaya. Earlier records include Ujung Kulon NP on the tip of Western Java, and Anakan Lagoon near Cilacap on the south coast. They have disappeared in most sites where they were known earlier, with the loss of their habitat to development and urbanization and with trapping. Their numbers are low and their conservation status is vulnerable (Andrew 1990, MacKinnon and Phillips 1993, BirdLife International 2001).

Habitat and general habits Mangrove and other wooded tidal swamps in estuaries, associated with swamp fern Acrostichium, thickets and tall grasses Saccharum and Imperata in coastal lowlands, and Nypa palm swamps in brackish water behind mangroves; they may also occur in inland grass swamps. In coastal wetlands they occur in fringing swamps with pioneer mangroves Avicennia.They are not forest birds and are replaced

in dense stands of mature mangroves Rhizophora and Bruguiera by Greater Coucals Centropus sinensis. Where coastal habitat has been converted to fish and shrimp ponds and agricultural lands, they are replaced by Lesser Coucal C. bengalensis and Greater Coucal C. sinensis. In the wet season Javan Coucals have been seen in Imperata grassfields and swamps and along partly flooded forest edge, and they were netted in a sugarcane plantation near Surabaya (Ir Darjono, 2004). In the dry season they forage as they walk near puddles and on dry marshland and in high grass meadows (Bartels 1915–1930 diaries and notebooks on Javan birds, in BirdLife International 2001).

Food Omnivorous, they take many insects, including grasshoppers, beetles, large moths, hairy and smooth caterpillars, cicadas, large bugs and dragonflies; also snails, slugs, centipedes, bird and insect eggs, frogs, tree snakes and rodents (Bartels 1915–1930 and Sody 1953, in BirdLife International 1989). In one study the coucals took frogs Rana, geckos Hemidactylus, water snakes, rat Rattus dragonflies, beetles, grasshoppers and a cicada; in total prey numbers 40% amphibians and reptiles, 2% mammals and 34% insects, other items were not identified (Arifin 1997 in BirdLife International 2001).

Breeding Eggs seen in March and June; in a recent study the nesting period was January to March (Hellebrekers and Hoogerwerf 1967, Arifin 1997 in BirdLife International 2001). One nest was in an Acrostichium fern in a clearing along a river, a loosely-built structure of fresh and dry fern leaves, with a nest lining of fresh fern and grass leaves; other nests were 3–6 m above the surface in mangroves (Bartels 1915–1930, in BirdLife International 2001). Eggs are white, 39 ⫻ 31 mm, weight 1.8 g, clutch 1–3 (Bartels, in Hellebrekers and Hoogerwerf 1967) to 3–5 (BirdLife International 2001). Incubation and nestling periods are unknown.

238 Greater Coucal Centropus sinensis

Greater Coucal Centropus sinensis (Stephens 1815) Polophilus sinensis Stephens 1815, in Shaw’s General Zoology 9, pt. 1, p. 51. (“China” ⫽ Ning Po) Other common names: Crow-Pheasant, Common Coucal, Hume’s Crow-Pheasant, Larkheeled Cuckoo, Brown Coucal (andamanensis). Polytypic. Seven subspecies. Centropus sinensis sinensis (Stephens, 1815); Centropus sinensis bubutus Horsfield, 1821; Centropus sinensis andamanensis “Tyeleri” Beavan, 1867; Centropus sinensis intermedius Hume, 1873; Centropus sinensis kangeangensis Vorderman, 1893; Centropus sinensis anonymus Stresemann, 1913; Centropus sinensis parroti Stresemann, 1913.

Description ADULT: Sexes alike, female larger; plumage black on the head, mantle and underparts, head glossed blue to purplish, shaft streaks above are black, wing chestnut, tail black glossed greenish, long and broad, under wing coverts black (some birds have thin white bars); iris brown to red, bill black, feet black. There is no seasonal alternation of plumages by molt. The plumage of the head, mantle and breast varies in gloss, with some birds glossy blueor purple-black and others dull sooty black. The dull birds sometimes have remnant juvenile plumage; however, some birds are in molt from the barred juvenile plumage into glossy plumage on the head and breast. JUVENILE: Head dull black with brown spots, nape black with whitish bars, back black with rufous bars, upper tail coverts black with buff bars, wing coverts, inner secondaries, and tips of outer flight feathers rufous barred and tipped blackish, tail black with narrow buff bars, underparts from throat to under tail coverts blackish barred dull white with yellowish shaft streaks; iris gray to brown, lower mandible partly pale. Birds molt into an unbarred black body plumage while retaining a few juvenile barred flight feathers (Stresemann 1913a, 1939, Whistler and Kinnear 1934, UMMZ).

NESTLING: At hatching, skin black with long white hair-like down, the down hanging forward in a fringe over the eyes and bill, center of belly pinkish; upper mandible black with pink edges and an egg tooth, gape yellow, iris brown, feet brownish gray (Shelford 1900, Inglis 1903, Roberts 1991, Natarajan 1997). SOURCES: AMNH, BMNH, CM, FMNH, RMNH, ROM, SMTD, UMMZ, USNM, ZSM.

Subspecies Centropus sinensis sinensis (Stephens, 1815); as above; Pakistan, India from Punjab and Kashmir east through Himalayas and Gangetic Plain to Assam, Nepal, the Bhutan foothills and southern China (Guangxi, Zhejiang, Fujian); Centropus sinensis parroti Stresemann, 1913; upper back black, wing of juvenile without bars; peninsular India from Bombay, Madhya Pradesh and Orissa south to Kerala; and Sri Lanka; Centropus sinensis intermedius Hume, 1873; smaller; Bangladesh, India (west Cachar), Burma north to Chin Hills, China (Yunnan, Hainan), Thailand, Indochina and northern Malay Peninsula; Centropus sinensis bubutus Horsfield, 1821; wing paler rufous, larger; southern Malay Peninsula, Sumatra, Nias, Mentawai Islands, Java, Bali, Borneo, western Philippines (Balabac, Cagayan Sulu and Palawan); Centropus sinensis anonymus Stresemann, 1913; shorter wings, wing darker brown than bubutus; southwestern Philippines (Basilan, Sulu Islands); Centropus sinensis andamanensis Beavan, 1867; two plumage phases, one pale with head and body buffy white, wing brown, tail whitish above and rufous below; the other dark with head and back brown, rump darker, wing brownish purple with tip darker brown, tail bronzy purple, belly dusky, wing lining gray buff, iris red to yellow; Andaman Is and neighboring islands of Burma (Table, Great Coco and Little Coco Is);

Greater Coucal Centropus sinensis 239 Centropus sinensis kangeangensisVorderman, 1893; two plumage phases, one pale with head and body buffy white, wing rufous and tail rufous gray; the other dark with head, back, tail and throat brownish gray and the breast mottled gray, larger than andamanensis; Kangean Is. The mainland subspecies are not very distinct. In the Bombay Natural History Society collection, C. s. intermedius are not consistently smaller, sex for sex, than C. s. sinensis (Biswas 1960). In the Malay Peninsula around 5–6°N, C. s. intermedius intergrade with bubutus (Wells 1999). Sumatra birds do not differ from Java birds C. s. bubutus ( Junge 1948). On the other hand, C. s. parroti is distinct from C. s. sinensis. The island forms C. s. andamanensis and C. s. kangeangensis, each sometimes called a distinct species, have adults in a pale plumage phase and a dark plumage phase, the dark phase buff to brown rather than black as in adult C. sinensis on the mainland and larger islands.

Measurements and weights Centropus sinensis sinensis; Assam (Palasbari, Garo Hills, Khasi Hills, Naga Hills): Wing, M (n ⫽ 9) 197–208 (203.0 ⫾ 3.9), F (n ⫽ 10) 202–228 (214.6 ⫾ 9.1); tail, M 230–256 (245.8 ⫾ 9.1), F 241–277 (262.9 ⫾ 11.3); bill, M 31–34 (32.8 ⫾ 1.1), F 34–38 (36.1 ⫾ 1.6); tarsus, M 53–58 (54.9 ⫾ 2.2), F 56–64 (60.1 ⫾ 2.7); hallux claw, M 22–30 (25.6 ⫾ 2.4), F 27–32 (28.8 ⫾ 1.8) (UMMZ); Centropus sinensis parroti; Madhya Pradesh: Wing, M (n ⫽ 5) 176–197 (186.0), F (n ⫽ 5) 194–237 (210.4); tail, M 248–273 (266.3), F 258–290 (280.0) (UMMZ); Deccan to southern Indian peninsula and Sri Lanka: Wing, M (n ⫽ 23) 177–196 (187.6 ⫾ 5.1), F (n ⫽ 22) 196–212 (201.6 ⫾ 4.4) (Stresemann 1913a); Centropus sinensis intermedius; Thailand:Wing, M (n ⫽ 6) 184–212 (199.5 ⫾ 9.5), F (n ⫽ 6) 190–226 (207.0 ⫾ 13.0); tail, M 242–288 (263.3 ⫾ 16.5), F 250–290 (274.0 ⫾ 14.9); bill, M 30–36 (32.2 ⫾ 2.3), F 33–39 (35.5 ⫾ 2.6); tarsus, M 53–60 (57.2 ⫾ 2.4), F 55–60 (58.0 ⫾ 2.5); hallux claw, M 16–26 (23.0 ⫾ 3.6), F 17–28 (24.0 ⫾ 3.8) (ZMUC);

Centropus sinensis bubutus: Wing, M (n ⫽ 20) 202–221 (211.8 ⫾ 5.2), F (n ⫽ 18) 211–239 (227.4 ⫾ 7.3) (Stresemann 1913a); Centropus sinensis anonymus: Wing, M (n ⫽ 3) 182–191 (186.3), F (n ⫽ 1) 199 (Stresemann 1913a); Centropus sinensis andamanensis: Wing, M (n ⫽ 13) 167–200 (180.0 ⫾ 8.9), F (n ⫽ 13) 177– 190 (182.9 ⫾ 4.9); tail, M 235–260 (242.6 ⫾ 7.1), F 226–253 (242.9 ⫾ 9.4); bill, M 30–36 (33.0 ⫾ 2.0), F 30–36 (31.8 ⫾ 1.8); tarsus, M 41–50 (46.0 ⫾ 3.0), F 44–53 (49.0 ⫾ 3.6); hallux claw, M 20–24 (21.5 ⫾ 1.4), F 20–24 (23.1 ⫾ 2.2) (AMNH, BMNH); Centropus sinensis kangeangensis: Wing, M (n ⫽ 4) 192–205 (200.8 ⫾ 6.0), F (n ⫽ 2) 211–222 (216.5) (Stresemann 1913a); the hallux claw is long and straight (c. 18 mm) as in other C. sinensis. Weight, Nepal and India: M (n ⫽ 1) 255, F (n ⫽ 1) 370 (USNM, ZSM); India, C. s. parroti: M (n ⫽ 2) 208–218 (213) (Saha and Dasgupta 1992); Kangean Is, C. s. kangeangensis: M (n ⫽ 2) 254–270 (262), F (n ⫽ 4) 275–380 (305.8) (ZMB). Wing formula, P5 ⫽ 4 ⬎ 6 ⫽ 3 ⬎ 7 ⫽ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 47–52 cm.A large black coucal with chestnut wings and a long, broad, black tail (except in island forms). Island coucals (Andamans, Kangean) are polymorphic, the body either buff or gray-brown, with chestnut wings. Tail is proportionately longer than in Short-toed Coucal C. rectunguis. Differs from the smaller Lesser Coucal C. bengalensis in its black not pale shaft streaks above and the uniform rufous chestnut of the wings. Differs from Green-billed Coucal C. chlororhynchos in Sri Lanka by its black bill, paler wings, higher-pitched and longer calls and its jungle habitat. Differs from the smaller Philippine Coucal C. viridis by the black shaft streaks both above and below. Juvenile differs from C. bengalensis on the head and back, which are barred (streaked in C. bengalensis), and on the underparts, dark gray with narrow whitish bars (pale and streaked in C. bengalensis).

240 Greater Coucal Centropus sinensis

Voice Call, deep hooting notes, “hoop, hoop, hoop, hoop”, 3–4 notes a second, slow, low and mellow, the notes at 0.3 kHz given about 3 per sec, the note lasting 0.1 sec.The full song runs down and up the scale, faster in the middle of the series, followed by more “hoop” notes. Both sexes give the call, and a pair often duets with their calls overlapping in time, the female at a lower pitch than the male. Taking into account the differences in pitch of males and females, the lower pitch of the larger forms, and the variation in completeness of the repeated calls and songs, the geographic forms of these coucals appear to have nearly the same calls and are similar on mainland Asia and in the Andaman Is, although southern Indian birds appear to differ from northern Indian birds (P. C. Rasmussen). Other calls, a “tok, tok,” a rapid double-note rattle “lotok, lotok . . .”, a harsh scold “skaaah”, a soft “meeaow” by the female, and an alarm or threat “k’wisss”. Females and juveniles call “tch-truu, tchtruu” when they see the male carrying food (Smythies 1960, King and Dickinson 1975, Lekagul and Round 1991, Roberts 1991, Natarajan 1997, Scharringa 1999, Wells 1999, Kennedy et al. 2000, Robson 2000a, Sheldon et al. 2001, Supari 2003).A common name of the bird in Burma is “Bok” after the sound of the call (Smythies 1940), and an interpretation of the call “bu-bu-bu” in Borneo is the source of the local name and the subspecies name bubutus (Smythies 1999).

Range and status Indian subcontinent from Pakistan (Punjab and Indus plains and into the perennially irrigated tracts of the NW Frontier District), India from Jammu along the base of the Himalayas to Assam, south to Bangladesh and Bihar; from the Gangetic plains south from Orissa throughout the peninsula; southern China (south of Yangtse and the coastal plains in Yunnan, S Zhejiang, Fujian, Guangdong, E Gwangxi and Hainan), mainland and island southeast Asia; Sri Lanka, Nicobar Is, Sumatra (including Nias and Siberut), Borneo (including N Natuna Is), Java, Bali (including Kangean I) and the southwest Philippines (Sulu Islands, Cagayan Sulu, Basilan Is,

Balabac, Palawan) (Riley 1938, Smythies 1981, van Marle and Voous 1988, Sibley and Monroe 1990, Cheng 1991, Dickinson et al. 1991, Roberts 1991, MacKinnon and Phillipps 1993, Grimmett et al. 1999, Robson 2000a, Thomas and Poole 2003). Resident. Common. On the continent their range has extended in recent years into northern India (Smetacek 1974). They live on many offshore islands. They appeared on Krakatau I and were more common than C. bengalensis on Sebesy I in 1919, 36 years after the volcanic activity of 1883 exterminated all land birds on these islands (Dammerman 1992, Thornton 1996). The form andamanensis on Andaman Is and neighboring islands of Burma (Table, Great Coco, Little Coco Islands) where it is the only coucal (Stresemann 1939, Ali and Ripley 1969, Ripley and Beehler 1989), is considered near-threatened because of restricted distribution (Collar et al. 1994). The birds are common within their range (Tikader 1984, Davidar et al. 1996).

Habitat and general habits Forest edge and disturbed habitats, secondary forest, logged-out country, overgrown plantations, banks of large rivers, tall grassland, thickets, bamboo, reedbeds, grasslands and scrub near cultivation, cocoa, gardens, paddyfields, cover near swamps, streams and lakes, and in villages with shade trees. Widespread, except in dense primary forest. A bird of the lowlands in open grass and scrub country, in India they are mainly in the plains; in the hills they are not often seen above 900

Greater Coucal Centropus sinensis 241 to 1200 m, rarely to 2100 m; in Sri Lanka they have extended their range into the highest hill country, in Malay Peninsula they are below 700 m, in Sumatra below 700–800 m, in Borneo in the lowlands and in grassy uplands on the Kelabit plateau, and in Sabah from sea level to 750 m (Jerdon 1862, Hume and Oates 1890, Baker 1934, Ali and Whistler 1937, van Marle and Voous 1988,Thewlis et al. 1996, Robson et al. 1998, Grimmett et al. 1999, Smythies 1999, Wells 1999, Sheldon et al. 2001). In the Andaman Is, they are in forest, sugarcane and padi plantation and mangrove swamp (Ali and Ripley 1969). Greater Coucals are less common than Lesser Coucals where their ranges overlap. Terrestrial, they stay under cover. They stalk, walk, hop and run in pursuit of prey; they walk into, through and out of thickets looking for food, creep through shrubs, prowl on muddy river banks near the water edge, sometimes flash wings forward to flush their prey. They get into trees where they clamber on inclined branches, ascend into canopy, hop from branch to branch, and glide down to undergrowth. Flight is slow and labored; the bird alternates flaps and glides (Ali 1953, Smythies 1986). Feeding techniques include probing in holes of tree trunks and bark, gliding after insects that fall to the ground while birds feed in trees, jumping and catching low-flying winged termites in the air, chasing grasshoppers and lizards on the ground, and hopping and flying while chasing snakes on the ground (Natarajan 1993). They roost in reedbeds (Roberts 1991). Sun-bathe in morning and after rain, in singles or in pairs. The young coucals disperse from their parents’ territory a few months after they fledge. Territory size of nesting pair, 0.9 to 7.2 ha (mean 3.8 ha) in southern India (Natarajan 1997).

Food Opportunistic predator, takes insects (caterpillars, roaches, grasshoppers, katydids, beetles, larvae of rhinoceros beetles and other grubs, dragonflies, ants); centipedes, millipedes, scorpions, spiders, crabs, large garden snails (Achatina snails are smashed, soft parts then extracted), slugs, earth-

worms, small mammals (mice, hedgehogs), snakes, lizards, frogs and toads, bird eggs and nestlings; fruits (including of oil palm) and seeds (Legge 1880, Ali and Whistler 1937, Ali and Ripley 1969, Smythies 1981, Natarajan 1993, Sody 1989, Roberts 1991, Wells 1999). The coucals search under dense herbage for snails, and they scavenge on dead carcasses. Around villages in Tamil Nadu, snails Helix vittata are the main food (Natarajan 1993).

Displays and breeding behavior Monogamous, the birds occur in pairs. Male struts and chases female on the ground, she advances with her tail depressed and her wings drooped and quivered, then gives a harsh call, and the pair copulates on the ground or in a tree (Ali and Ripley 1969).The nest is built mainly by the male but the female also builds; the pair takes three to eight or more days to complete the nest (Dhindsa and Toor 1981, Roberts 1991, Natarajan 1997). The male brings food to the female before mating, and he feeds her during copulation or eats it himself after they mate.

Breeding In the rains; in northern India they nest mainly from June to September, in peninsular India from November to May (Hume and Oates 1890, Baker 1934, Ali and Whistler 1937, Ali and Ripley 1969, Gaston 1981, Zacharias and Gaston 1983, Natarajan 1997), in the Andamans mainly at the end of the hot season and beginning of the rains, May to July, and in February and April (Hume 1874, Hume and Oates 1890, Baker 1934), in Sri Lanka nearly all year with a peak in March and April (Henry 1971), in southeast Asia from January to August (Robson 2000a), in Burma from April to August (Oates 1877, Hume and Oates 1890, Baker 1934) and a nest fledged in October (Stresemann and Heinrich 1939), in Thailand from May to September and a half-grown young was taken in February (Riley 1938), in the Malay Peninsula eggs occur from January to May (Wells 1999), in Java from January to April, June and October to December (Hoogerwerf 1949, Hellebrekers and Hoogerwerf

242 Goliath Coucal Centropus goliath 1967), in Sabah nests with eggs are seen in September and March and nests with young in November, December, February and March (Sheldon et al. 2001). The nest is a large globular dome of twigs and leaves, coarse grass or reeds, palm leaves or coconut frond fibers or the ribbon-like saw-toothed leaves of screw-pine Pandanus, and it is sometimes strengthened with mud. It is lined with leaves and grass, and it has a side entrance. Some nests have the sides thick, the back thin, and the bird sits with its head towards the back of the nest and the long tail hanging out the side entrance (Herbert 1924). Less often the nest is open, built of green leaves and lined with dry grass (Legge 1880, Macdonald 1906).The nest is built under cover of creepers and vines, concealed in thick bush, fern tangles and pandanus crowns, as high as 6 m above ground or low in a thorny tree, or on the ground in dense clumps of grass or in rice fields (Baker 1934, Ali and Ripley 1969, Natarajan 1997, Wells 1999, Sheldon et al. 2001). Eggs are chalky white, with a yellowish glaze when laid; this wears away leaving a white chalky shell, often stained by nest dirt (Hume and Oates 1890); size 36 ⫻ 28 mm and weight 14.8 g in India (Baker 1934), 34 ⫻ 29 mm in Sri Lanka (Henry 1971), 35 ⫻ 28 in the

Andaman Is (Baker 1934), 38 ⫻ 30 mm in Java (Hoogerwerf 1964), 33 ⫻ 28 in Sabah (Sheldon et al. 2001); smaller eggs of 29 ⫻ 24 mm in Borneo (Gibson-Hill 1949c) appear to be of the smaller Lesser Coucal C. bengalensis. Clutch 1–4 (mean, 2.7) in India, 3(4) in Pakistan, 2–5 in southeast Asia, 2–3 in Sri Lanka, 3–4 (once, 5) in Thailand, 2 in the Malay Peninsula, 3 in Sabah and 2–3(4) in the Andamans (Herbert 1924, Baker 1934, Riley 1938, Ali and Ripley 1969, Roberts 1991, Wells 1999, Robson 2000a, Sheldon et al. 2001). Eggs are laid before the nest is complete. Incubation period is 15–16 days. Both sexes incubate the clutch and feed the brood. During the first four days the adults provide nestlings with regurgitated food, mainly snails. Parents visit the nest 2–5 times an hour depending on the size and number of nestlings. Both parents remove fecal sacs. Nestlings are half the fledging weight by 8 days, reach fledging weight by 16 days, and fledge at 160 g in 18–22 days, then accompany and beg from their parents for another 2 months. In a population in India, 77% of the eggs hatched and 67% fledged. Nests with eggs were sometimes deserted, and many nests were robbed by Jungle Crow Corvus macrorhynchos (Natarajan 1997).

Goliath Coucal Centropus goliath (Bonaparte 1850) Centropus goliath Bonaparte 1850, “Forsten” Conspectus Systematis Ornithologiae, 1, p. 108. (Halmahera) Other common names: Giant Coucal. Monotypic.

Description ADULT: Sexes alike, upperparts and underparts black, wing coverts black with a large white or whitish patch, wing glossed bluish black, tail long, graduated, broad and black; iris dark brown, bill black, feet black. On Halmahera there is a whitish phase with buff head, and some birds have pied

plumage or have irregular white marks elsewhere in the plumage. JUVENILE: Head, neck and underparts from throat to breast blackish streaked whitish, upper back and lesser wing coverts with broad whitish diamondshaped spots, feather shaft whitish, barbs next to the shaft buff and white and dark brown around the edge of the feather, wing coverts black with a large white patch, wing black, secondaries and coverts often have whitish diamond-shaped streaks, tail glossed bluish black, face brown with indistinct whitish streaks, chin whitish, breast and belly blackish or (on Halmahera) irregularly blotched

Goliath Coucal Centropus goliath 243 whitish and gray or blackish and gray; bill dark along culmen and the rest of bill pale, feet pale. NESTLING: Head with long white to buff hair-like down; iris dark gray, bill black above and pale gray below, feet pale (ZMA 48.175). SOURCES: AMNH, BMNH, MCZ, MSNG, MZB, RMNH, ROM, UMMZ, USNM, YPM, ZMA.

Measurements Wing, M (n ⫽ 10) 244–290 (261.6 ⫾ 14.6), F (n ⫽ 8) 255–298 (274.9 ⫾ 13.8); tail, M 390–434 (421.6 ⫾ 22.1), F 360–425 (394.5 ⫾ 26.6); bill, M 48.5–58 (50.6 ⫾ 4.0), F 48–60.4 (55.0 ⫾ 5.5); tarsus, M 46.5–55.5 (52.4 ⫾ 3.2), F 48–57 (55.1 ⫾ 4.0); hallux claw, M 22.5–28 (25.3 ⫾ 2.4), F 21–27 (24.3 ⫾ 2.2) (AMNH, USNM). Weight, M (n ⫽ 7) 340–453 (400.7), F (n ⫽ 2) 605, 638 (621.5) (MZB, ZMA). Wing formula: P 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 7 ⫽ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 62–70 cm. Very large coucal with black plumage, a large white wing patch, and a long graduated tail.

Voice Advertising call, a persistent series of deep “ooom” notes repeated about 2 / sec, the series lasting 4–12 sec. In alarm, one or more birds give a series of harsh, guttural monosyllabic or disyllabic notes, “kcau” or “kcau-kuc,” repeated at 1–4 sec intervals. Other calls include a deep, low, moaning roar (Ripley 1959, Coates and Bishop 1997).Voice is lower in pitch than C. sinensis and C. celebensis (Heinrich 1956).

Range and status Northern Moluccas (Morotai, Halmahera, Ternate, Tidore, Kasiruta, Bacan, Obi). Resident. Locally common and conspicuous on Halmahera, uncommon on Bacan (Shelley 1891, Lambert and Young 1989, Coates and Bishop 1997), not seen recently on Obi or Ternate (White and Bruce 1986, Lambert 1994). They perch on shaded branches in midstory

or lower crown canopy at forest edge where they give far-carrying calls (Coates and Bishop 1997). Tobelo people of Halmahera have different names for the two plumage morphs, “o ciungu” for the bird in normal black plumage and “o ciungu ma heheke” (ugly, dirty) for the whitish morph (Taylor 1990). Although they have been considered an allospecies with Greater Black Coucal C. menbeki of New Guinea (White and Bruce 1986), they are more closely related to Pheasant Coucal C. phasianinus of New Guinea and Australia.

Habitat and general habits Forests, including larger patches of scrub at edge of agricultural land; sea level to 450 m on Halmahera (Heinrich 1956), in primary and logged forest to 250 m on Bacan (Lambert 1994). They live in undergrowth to midstory; occasionally they perch in higher trees. Clamber slowly up vine- and epiphyte-covered trees. They feed in the understory and in lower forest subcanopy to 12 m, and they glean from tree trunks, the bases of Livistona palm fronds, vine tangles and epiphytes. Usually seen in pairs or small family groups of 3–4 birds (Rand and Gilliard 1967, Coates and Bishop 1997).

Food Large insects, including crickets, grasshoppers, cicadas and phasmid walking-sticks (Rand and Gilliard 1967).

Breeding Unknown.

244 Madagascar Coucal Centropus toulou

Madagascar Coucal Centropus toulou (P. L. S. Müller, 1776) Cuculus Toulou P. L. S. Müller, 1776, Des Ritters C. von Linne . . . vollständiges Natursystems . . . Supplement, 90. (Madagascar) Other common names: Malagasy Coucal. Polytypic. Two subspecies. Centropus toulou toulou (Müller, 1776); Centropus toulou insularis Ridgway, 1894, Aldabra I; Assumption I (extinct).

Description ADULT: Sexes similar. Breeding plumage, head and upper back glossy black, hackles black, wing unbarred rufous chestnut with dark brown tips, inner secondaries blackish chestnut, back, rump and upper tail coverts unbarred black, tail glossy black; underparts, chin to breast black, marginal under wing coverts black and gray, greater under wing coverts rufous chestnut; iris red, bill black, feet gray to black. Nonbreeding plumage dark and streaked, upperparts black finely streaked with long pale yellow shafts, the inner part of barbs graybrown, wing rufous, unbarred; underparts, throat and breast blackish streaked with yellow and gray, belly and under tail coverts black; bill rose-brown. Some birds retain a few unbarred juvenile rectrices in breeding plumage, and these may breed at a year of age. JUVENILE: Crown black with inconspicuous rufous streaks and yellow shafts, back black with indistinct buff bars and short pale shaft streaks, the buff marks forming short triangles; wing rufous chestnut with dark brown bars and tips, rump blackish barred with buff, upper tail coverts sometimes long (80 mm), black with narrow buff bars (bars of 3–4 mm black, 1.5 mm buff ), tail black with buff bars and tip; underparts, chin and throat variable, black with small buff spots to buff with irregular black bars, breast and belly sooty gray, feathers with indistinct black subterminal bars, rachis orange in the gray area and black in the barred area of the feather, under wing coverts dark gray with paler bars.

NESTLING: At hatching skin is black, dorsal surface covered with dense white hair-like shafts; bill black with white dorsal tip (remnant egg tooth), mouth lining and palate pink, papillated edge of choana white, base of tongue whitish with a black Ushaped band near the tip (Appert 1980). SOURCES: AMNH, BMNH, FMNH, MNHN, MSNG, ROM.

Subspecies Centropus toulou toulou (Müller, 1776); as above; Madagascar; Centropus toulou insularis Ridgway, 1894; underparts paler in nonbreeding plumage; Aldabra, formerly Assumption I where it now is extinct. Although it is paler on average, the nonbreeding plumage is not consistently paler than in Madagascar.

Measurements and weights Wing, M (n ⫽ 7) 135–156 (148.0 ⫾ 6.1), F (n ⫽ 13) 156–176 (167.8 ⫾ 6.5); tail, M 198–248 (218.0 ⫾ 21.8), F 218–270 (241.4 ⫾ 33.1); bill, M 24–28 (25.7 ⫾ 1.2), F 28–31.6 (29.5 ⫾ 1.4); tarsus, M 34–38 (35.5 ⫾ 1.6), F 35–42.7 (38.5 ⫾ 2.2); hallux claw, M 17.3–24 (20.8 ⫾ 2.0) F 20–24 (21.6 ⫾ 1.4) (AMNH, FMNH). Weight, Madagascar, M (n ⫽ 4) 125–150 (135), F (n ⫽ 3) 185–220 (205) (Goodman and Benstead 2003); Aldabra, M (n ⫽ 1) 117, F (n ⫽ 1) 131 (Benson et al. 1976). Wing formula, P 5 ⱖ 6 ⬎ 4 ⬎ 3 ⬎ 2 ⬎1 ⬎7 ⬎ 8⬎9⬎10.

Field characters Overall length 40 cm (male), 46 cm (female). Coucal with black head and underparts, rufous wings in the breeding season, back and breast finely streaked rufous and black in nonbreeding plumage, and with a long tail. The juvenile is dark and barred.This is the only coucal on Madagascar.

Madagascar Coucal Centropus toulou 245

Voice Muffled series of five to ten hoots decreasing in volume, short notes given c. 6 per sec at 0.6 kHz; the hoots are a self-naming “toulou, toulou . . .”. Also gives water-bottle call duets, the first bird in a “toulou” series, the second joining with a higherpitched series at 14 notes per sec dropping in pitch and slowing to 8–10 notes per sec; another call is a sudden guttural “coogoo” (Rand 1936, Benson and Penny 1971, Roché 1971, Milon et al. 1973, Frith 1975, Langrand 1990, Morris and Hawkins 1998, Randrianary et al. 1997).

Range and status Madagascar and Aldabra. Resident.Widespread and not uncommon on Madagascar, except on the denuded central plateau. Population is estimated at several hundred on Aldabra, where the birds are rather tame around humans (Benson and Penny 1971, Betts 2002).The population on Assumption I became extinct owing to disruption with the mining of bird guano and the destruction of the soil and vegetation (Stoddart 1984).

Habitat and general habits Dense vegetation or underbrush in forest, primary forest, recent clearings, second growth and scrub, eucalyptus woodlands, littoral forest, palms, gallery forest, mangroves, marshy reedbeds and grass, rice paddies, gardens. Occur from sea level to 1800 m; at PN de Marojejy from 450 to 1250 m (Goodman et al. 2000).They stay low in vegetation, feed on the

ground and in thick scrub.They chase grasshoppers and lizards in a “flush and rush” behavior and they rip bark from trees to get lizards beneath the bark. Coucals often feed in pairs, when one bird defends the other against theft of food by drongos (PrysJones and Diamond 1984). They are mobbed by small passerines. The coucals seldom fly, and when they do they give vigorous flaps, then glide into vegetation.

Food Invertebrates, large insects (beetles (Elateridae, Scarabaeidae, Tenebrionidae, Tettigoniidae), roaches (Blattodea), grasshoppers, crickets, mantids (Mantidae), caterpillars), ants (Formicidae), spiders (Salticidae), centipedes (Scolopendromorpha), mollusks, lizards (geckos, skinks), rats, chicks and eggs (Milon et al. 1973, Frith 1975, Goodman et al. 1997).

Displays and breeding behavior Male feeds the female with an insect in courtship. The pair wag their tails from side to side, calling occasionally; the male mounts as the female lowers and extends her wings, then passes the insect to her (Frith 1975).

Breeding Season uncertain. Adults are seen in black breeding plumage from July to September (Bangs 1918), and there are nesting records from September to March (Rand 1936, Milon et al. 1973, Frith 1975,Woodell 1976a, Benson et al. 1976, Appert 1980, Goodman et al. 1997, FMNH). Frith’s (1975) report that birds sometime breed in the brown “non-breeding” plumage on Aldabra suggests that the breeding season cannot be determined from the plumage alone, although other observations are of birds breeding only in the black plumage (Prys-Jones and Diamond 1984, R. Prys-Jones pers. comm.). In contrast to other birds but like some other coucals, the right testis is larger than the left testis (which is small in most coucals, and is sometimes absent) (Rand 1933). The nest is either a large spherical dome or a bulky open platform, woven from dry grass, twigs, creepers and other plant material longer than 10 cm, and lined with fresh green leaves. It is built

246 African Black Coucal Centropus grillii by the male, above the ground near the stem of a tree in dense vegetation such as a liana (Rand 1936). Eggs are white, 33 ⫻ 2 6 mm (Rand 1936). An egg is sometimes laid before the nest roof is built. Clutch size in Madagascar is 2–4, in Aldabra 2. Some eggs are laid several days apart. The male incubates more than the female, and he attends the nest. Incubation period is 14–16 days, the young hatch asynchronously when eggs are laid at intervals as long as 9 days apart (or as short as 3 days

apart, Woodell 1976b)—they hatch at intervals as long as 7–8 days apart.The parent feeds crickets and grasshoppers to the nestlings. The young beg with a churring sound.When disturbed in the nest they give a high-pitched hiss and excrete a foul-smelling black, sticky fluid. Nestlings grow to c. 90% of fledging weight by 15 days.The flight feathers and tail feathers are nearly half grown and most contour feathers open by 16 days, and the young fledge in 19 days (Frith 1975).

African Black Coucal Centropus grillii Hartlaub, 1861 Centropus Grillii Hartlaub, 1861, Journal für Ornithologie, 9, p. 13. (Gabon) Other common names: Black Coucal, Blackbellied Coucal Monotypic.

Description ADULT: Sexes alike. Breeding plumage, head and body black, black shaft streaks on the head, upper back and breast, lower back black barred buff, wing coverts rufous brown, wing rufous, tips of primaries and outer secondaries dark, inner secondaries dark brown, tail black; iris brown, bill black, legs and feet black. In nonbreeding plumage, upperparts dark brown with rufous barring, forehead to upper back streaked light rufous and black, wing rufous, tail black, no eye streak; underparts buff with whitish shaft streaks, belly with narrow brown bars, under wing coverts dark rufous.After the breeding season, black-plumaged adults molt into brown plumage and the bill becomes horn-gray. Birds in their first breeding season sometimes retain the barred wing coverts and flight feathers of the juvenile plumage in the black adult breeding plumage. JUVENILE: Upperparts light rufous and dark brown, head streaked and mottled with pale buff triangles on tip of feathers, nape and upper back streaked whitish along shaft, rufous along the white and dark brown along the feather edge, back barred light rufous and dark brown, wing coverts and

flight feathers rufous with dark brown bars, rump and upper tail-coverts dark brown with narrow buff bars, tail dark brown with fine rufous bars (dark bars 8 mm wide, buff bars 2 mm wide), under wing coverts pale chestnut with fine black bars; underparts, cheeks whitish, chin and throat whitish with fine black streaks, breast whitish buff with paler shaft streaks and barred black on the flanks, belly whitish, lower belly gray and under tail coverts blackish with buff bars; iris pale gray, bill pale yellowish, feet blue gray. NESTLING: At hatching, skin black with long white hair-like down above, forming a fringe over the forehead; bill black with a white egg tooth. SOURCES: AMNH, BMNH, BWYO, CM, FMNH, KMMA, NMM, ROM, SMF, USNM, ZMUC, ZSM.

Measurements and weights Wing, M (n ⫽ 14) 146–158 (152.3 ⫾ 3.9), F (n ⫽ 14) 164–173 (169.1 ⫾ 3.5); tail, M 122–170 (149.2 ⫾ 14.0), F 132–170 (159.0 ⫾ 10.1); bill, M 19–25 (22.29 ⫾ 1.6), F 22–27 (24.36 ⫾ 1.6); tarsus, M 29–40 (34.6 ⫾ 3.8), F 35–42 (38.9 ⫾ 2.5); hallux claw, M (24.6 ⫾ 3.6), F 23–31 (27.8 ⫾ 2.1) (BMNH (excludes specimens of A. Whyte, perhaps missexed), FMNH, ZSM). Weight, M (n ⫽ 6) 94–108 (100), F (n ⫽ 1) 151 (Irwin 1988). Wing formula, P 7 ⬎ 6 ⬎ 5 ⬎ 8 ⫽ 4 ⬎ 3 ⬎ 9 ⬎ 2 ⬎ 1 ⬎ 10.

African Black Coucal Centropus grillii 247

Field characters Overall length 30 cm (male), 34 cm (female). Small coucal in marshes and wet grass, breeding plumage black with rufous wings. Non-breeding and juvenile birds are streaked and barred, juveniles have barred wing and tail.The tail is shorter than in other coucals. They lack an eye streak (present in some plumages of White-browed Coucal C. superciliosus).

X

X

Voice The call usually heard is a double note “kop-kop” repeated at 2-sec intervals. Long song is a series of “kop” notes each rising and falling in pitch in the range 0.8–1.0 kHz, repeated 6 notes per sec for 2–3 sec with no change in pitch. When I watched the coucals at Ngaoundéré, Cameroon, in October, the two-note call was given in the early morning as the bird held its head down, with neck puffed out and bowed to the breast, a posture that other coucals use when they give the long water bottle call. Excitement or aggressive call is a series of harsh “shrehhh” and alarm call is a slow clucking “tucktuck” (Chappuis 1974, 2000, Stjernstedt 1993).

Range and status Africa from Senegal,The Gambia, Liberia and Ivory Coast eastward to northern Cameroon, Central Africa from Gabon and Angola through S and E Zaire to Ethiopia, Uganda and Kenya, also Zambia, Malawi, northern Zimbabwe, N and E Botswana, Mozambique and NE Natal. Resident or seasonal, depending on whether large regions or local sites are indicated. They appear and disappear with the rains, migrating into low-rainfall grasslands in the rainy season.They occur in both wet and dry seasons on the upper Niger River in Mali (Lamarche 1980). Seasonal with the rains from August to October in The Gambia; some remain into December (Barlow et al. 1997, Barlow 2003). In Ghana they occur in the south from May to August and in the north in the wet season from July to September (Greig-Smith 1976, Grimes 1987). In Togo they are seasonal, in the humid south in the dry season from November to July, in the north in breeding plumage in July to August (Cheke and Walsh 1996), and in southern Benin they occur from January to June

(Anciaux 2000). Resident in southern Nigeria and seasonal from April to October during the rains in northern Nigeria (Bannerman 1951, Elgood et al. 1994). In Gabon they occur in the dry months January and February (Rand et al. 1959, Brosset and Erard 1986). In Sudan and Ethiopia they are resident (Urban and Brown 1981, Nikolaus 1987), in East Africa they are seasonal with the rains (Lewis and Pomeroy 1989). On the Kafue River floodplain of Zambia they arrive in December or January and leave in April or May after breeding there (Osborne 1973). In the Okavango they may be resident (Skinner 1996) and in eastern Botswana they appear with heavy rains (Brewster 2000); in Zimbabwe they occur mainly from December to April (Irwin 1981). In South Africa they are uncommon irregular seasonal migrants in summer in Kruger NP and Natal (Vernon et al. 1997). The coucals are accidental on Zanzibar during the mainland dry season (Pakenham 1979). Common in suitable habitat, scarce over the total range. Seasonal disappearance of surface waters after rains and through the dry season and the largescale conversion of wetlands into sugarcane and the burning of old grasslands for grazing, all affect the local status of these coucals.

Habitat and general habits Tall rank grass and reedbeds, grassland near freshwater swamps in wide grassy river valleys and marshes, in floodplains and seasonally flooded grasslands (Irwin 1988, Lewis and Pomeroy 1989, Leonard 1998). In West Africa, they occur in high rank grass, in dry rice fields and forest clearings and coastal

248 Philippine Coucal Centropus viridis airstrips (Thiollay 1985, Gatter 1997, Barlow 2003). The coucals remain low in the grass under cover when not calling. Feeding behavior has not been seen as the bird creeps through the grass. Seen when it flies over the grass, alternating flaps and glides and flopping into cover (Winterbottom 1938).

Food Insects—mainly grasshoppers, also beetles, caterpillars, hemiptera, crickets, mantids, ants; spiders, small reptiles, and seeds (Chapin 1939, Irwin 1988).

Displays and breeding behavior Territorial, the pair defends its area against neighbors (Winterbottom 1938), the area 5–10 ha (Vernon 1971). In courtship, the female takes the initiative. Both sexes silently quiver their lowered wings before mating, male feeds female at copulation. Females are sometimes polyandrous (once 1 female observed with 3 males, each male tended a nest) and lay several clutches in a season (Vernon 1971, Rowan 1983); more often they are monogamous with two adults tending the young (Barlow 2003).

Breeding During rains when the grass is high in The Gambia, fledged young with short tails and a yellow gape were tended by two adults in eclipse plumage in December (Barlow 2003). Birds are in breeding plumage in Liberia from February to August (Gatter 1997), in northern Ghana birds in July and August, in coastal Ghana from April to July (Holman 1947,

Greig-Smith 1977, Grimes 1987), in northern Nigeria breeds in July and August (Elgood et al. 1994), in northern Tanzania (Serengeti, Mkomazi) from December to February (Brown and Britton 1980, Zimmerman et al. 1996), in Zambia from November to April (Aspinwall and Beel 1998), in Malawi from January to May (Benson and Benson 1977), in Botswana in January and February (Skinner 1996), and in Zimbabwe from December to April (Vincent 1946, Irwin 1981, 1988, Vernon 1971, Vernon et al. 1997). The nest is an oval ball of dry grass and sedge, concealed in grass or sedge 20–40 cm above the ground, with growing stems of grass wrapped around and over the nest chamber to form the outer covering, while the lower part of the nest is solid, substantial and lined with leaves ( Vincent 1946). The nest is lined with green leaves which are added before laying and during incubation. Eggs are white, 31 ⫻ 24 mm, clutch 3–6, eggs laid at irregular intervals (every day or with gaps up to 9 days). Incubation begins with the first egg.The incubation period is about 14 days. The nestling fledges when undisturbed at 18–20 days, or earlier when the nest is disturbed, then returns to the nest; the young bird flies at 28 days.The adult male builds the nest, incubates and provides all the care to the young, making 6–9 visits per hour to the nest, mainly in the early morning while the female keeps track of the nests and replaces clutches that are lost (Vernon 1971, Irwin 1988). Breeding success-a group of 4 adults reared c. 19 young in 6 breeding attempts (Vernon 1971).

Philippine Coucal Centropus viridis (Scopoli, 1786) Cuculus viridis Scopoli, 1786, Delicae Florae et Faunae Insubricae, 2, p. 89. (Antigua, Panay [Philippines]) Polytypic. Four subspecies. Centropus viridis viridis (Scopoli, 1786); Centropus viridis mindoroensis (Steere, 1890); Centropus viridis carpenteri Mearns, 1907; Centropus viridis major Parkes and Niles, 1988.

rufous, wing coverts rufous, wing rufous with dark brown tips, tail black, underparts with pale buff shaft streaks, under wing coverts dark gray with narrow whitish bars; other geographic forms are all black, though there is an uncommon white phase on Luzon; bare skin around eye gray, iris red to brown, bill gray to black, tarsus dark gray.

Description ADULT: Sexes alike, head, back and underparts black glossed blue to green, hackles with black shaft streaks from head to back and breast, lower back

JUVENILE: Above grayish brown streaked whitish buff, the whitish shaft streaks widening to elongate whitish spots on the upper back, back and rump

Philippine Coucal Centropus viridis 249 chestnut barred rufous, wing barred grayish brown and rufous, tail long, unbarred bronze green (tip sometimes barred), rectrices narrow; underparts irregularly barred whitish buff and dark gray, paler on throat and darker on belly, throat and upper breast sometimes streaked with whitish shaft streaks, under tail coverts slate gray, tail and body plumage soft texture. NESTLING: “Naked” at hatching, skin dark (Rabor 1977). SOURCES: AMNH, BMNH, CM, CMNH, FMNH, MVZ, ROM, SMTD, UMMZ, USNM, ZMUC.

Subspecies Centropus viridis viridis (Scopoli, 1786); wing chestnut, smaller; the Philippines (including Luzon, Mindanao, Masbate, Bohol, Negros, Cebu, Leyte, Samar, Catanduanes); Centropus viridis major Parkes and Niles, 1988; wing chestnut, larger; Babuyanes Islands group (the Philippines); Centropus viridis carpenteri Mearns, 1907; wing black ( juvenile and adult), larger; Batanes Islands group (the Philippines); Centropus viridis mindoroensis (Steere, 1890); wing black, smaller; Mindoro and Semirara (the Philippines). Comments: Ross and Ramos (1992) found a small bird on Camiguin (wing 162, tail 175), like C. v. viridis on Luzon rather than large like C. v. major on the other Babuyanes islands. Subspecies ranges here are as summarized by Parkes and Niles (1988), Dickinson et al. (1991) and Ross and Ramos (1992).

Measurements and weights C. v. viridis; Luzon: Wing, M (n ⫽ 8) 142.5–155.5 (148.4 ⫾ 4.7 66), F (n ⫽ 8) 151–178 (159.9 ⫾ 8.5); tail, M 223–234 (228.2 ⫾ 5.6), F 235–265 (247.6 ⫾ 9.8); bill, M 26.9–29.2 (28.1 ⫾ 1.0), F 227.6–31.5 (29.8 ⫾ 2.2); tarsus, M 35.5–38.5 (37.0 ⫾ 0.9), F 37.3–42 (40.0 ⫾ 1.8); hallux claw, M 16.6–21.2 (20.7 ⫾ 1.7), F 21.7–28.2 (23.5 ⫾ 3.2) (FMNH);

C. v. major: Wing, M (n ⫽ 5) 163–182 (171.0), F (n ⫽ 5) 175–190.5 (182.6) (Parkes and Niles 1988, Ross and Ramos 1992); C. v. mindoroensis: Wing, M (n ⫽ 9) 143–158 (153.7), F (n ⫽ 5) 163–175 (171.8) (Ripley and Rabor 1958); C. v. carpenteri: Wing, M (n ⫽ 1) 173, F (n ⫽ 4) 181.5–195 (189.5) (Parkes and Niles 1988, SMTD). Weight, C. v. viridis; Luzon: M (n ⫽ 8) 100.2–126.7 (112.1), F (n ⫽ 8) 108.4–152.1 (133.5) (FMNH); C. v. major: M (n ⫽ 2) 155–169 (162), F (n ⫽ 3) 212–223 (218) (Ross and Ramos 1992); C. v. carpenteri: M (n ⫽ 2) 153–170 (161.5), F (n ⫽ 6) 179–253 (209) (CMNH, USNM); C. v. mindoroensis: M (n ⫽ 6) 104–140 (123), F (n ⫽ 2) 150–165 (157.5) (Ripley and Rabor 1958). Wing formula, P5 ⬎ 6 ⬎ 4 ⬎ 3 ⬎ 7 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 41–43 cm. Coucal with black head and body, and chestnut wings with dark brown tips. Birds on Mindoro and Batanes Islands are all black including the wings, the only all-black coucal in the Philippines (the larger C. steerii on Mindoro is dark brown).The birds are larger than the shortertailed Lesser Coucal C. bengalensis.

Voice A long, loud “boop boop boop . . .” given day and night, the notes about 0.6 kHz in pitch and given about 6 notes per sec; also a faster staccato “coo-coo-coo . . .” often descending in pitch, with 10 notes per sec, and “cha-coo” and “chi-go-go gook” alarm calls (Rabor 1977, Scharringa 1999, Kennedy et al. 2000).

Range and status Philippines, widespread except on Palawan and the Sulu Is, where Greater Coucal C. sinensis occurs. Resident. Common, it is the most common coucal in the Philippines (Stresemann 1939, Gonzales 1983, Dickinson et al. 1991).

Habitat and general habits Tall dense grass, mixed cultivation, disturbed second growth, thickets and vine tangles, bamboo, open

250 Lesser Coucal Centropus bengalensis

Food Insects, including beetles, grubs, grasshoppers and caterpillars, spiders, small lizards, sometimes carrion (Gonzales 1983, Goodman and Gonzales 1990, ZMUC).

Breeding

grass, clearings and edges of cultivation and gardens with dense vegetation; lowlands to mountains as high as 2000 m, on Mt Isarog uncommon at 450 m (Rabor 1977, Gonzales 1983, Goodman and Gonzales 1990, Kennedy et al. 2000, Fisher and Hicks 2000). Terrestrial, they live on or near the ground. When flushed out they rise and fly a short distance then plunge into thick bush or grass (Wolfe 1938, Gilliard 1950, Rabor 1977). Usually feed on or near the ground, hop under cover and climb through thickets of thorns and branches, seen in early morning or after a rain perched high while sunning and preening feathers to dry (Gonzales 1983).

Prolonged season; eggs are laid in late February (to account for an immature in April, Rand and Rabor 1960), a female had 3 ovulated follicles in April (CMNH 36135); other dates extend from May to July (Hachisuka 1934, Rabor 1977, Dickinson et al. 1991, Goodman et al. 1995, Robson 2000b, SMTD), and a large nestling was taken in September (ZMUC). Nest is a bulky globe of grass, about 20 cm h i g h ⫻ 13 cm wide or larger, built from wide grass blades woven around standing grass stems and incorporating live grasses pulled into the sides and over the top; entrance on one side; in tall grass usually 0.3–1 m above ground, lined with green grass and green leaves, which are added after the eggs are laid (Wolfe 1938). Eggs are dull white, chalky, 29.4 ⫻ 24.7 mm (Wolfe 1938), 30.4 ⫻ 25.0 mm (Rabor 1977), the clutch 2–3 (Wolfe 1938, Rabor 1977, CMNH 36135).The incubation period is about 2 weeks (Rabor 1977).The nestling period is unknown.

Lesser Coucal Centropus bengalensis (Gmelin, 1788) Cuculus bengalensis Gmelin, 1788, Systema Naturae 1, pt. 1, p. 12. (Bengal) Polytypic. Five subspecies. Centropus bengalensis bengalensis (Gmelin, 1788); Centropus bengalensis javanensis (Dumont, 1818); Centropus bengalensis medius Bonaparte, 1850; Centropus bengalensis lignator Swinhoe, 1861; Centropus bengalensis sarasinorum Stresemann, 1912.

Description ADULT: Sexes alike, breeding plumage, head and upper back black with green to blue gloss and black shaft streaks, lower back and rump brown, upper tail coverts black, tail glossy black sometimes tipped

whitish buff, wing coverts rufous brown with fine whitish shaft streaks, wing rufous brown with dark brown tips, underparts black from chin to under tail coverts, wing lining chestnut brown; iris red to reddish brown, bill black, legs black. Adult nonbreeding plumage is rufous brown with pale streaks. Head and upper back are brown with feathers having whitish streaks along the shaft, dark brown on the middle of the barbs and rufous on the edges; back unbarred brown; central upper tail coverts are barred and unusually long, in some birds nearly as long as the longest rectrices T1. Wing unbarred rufous brown with dark brown tips, tail

Lesser Coucal Centropus bengalensis 251 unbarred black or black with buff bars. Underparts buffy with fine straw-colored shaft streaks and indistinct incomplete black bars on throat and breast to belly, center of belly unmarked whitish, and flanks, lower belly and under tail coverts barred with black; under wing coverts pale chestnut; bill pale. Field observations indicate a seasonal change from black plumage to streaked brown plumage in India, southeast Asia, the Philippines and Taiwan (Deignan 1955, King and Dickinson 1975, Rabor 1977, Severinghaus and Blackshaw 1976, Grimmett et al. 1998); no museum specimens are in this molting plumage (Stresemann 1913b, Mees 1971). First-year nonbreeding (winter) plumage, head and back streaked brown, wing coverts and flight feathers rufous with blackish bars, central tail coverts barred and nearly as long as the tail, tail barred buff and dark brown. Underparts are as in older adult nonbreeding plumage, whitish to rich buff with streaks, dusky spots and bars, and under wing coverts pale chestnut. First winter plumage is attained by an incomplete molt of juvenile plumage, retains some barred feathers of the wing and tail; complete molt of juvenile body plumage. Birds molt directly from nonbreeding first-year plumage into adult breeding plumage, and some birds retain a few barred wing and tail feathers into the first breeding plumage. JUVENILE: Upperparts light rufous and dark brown, head irregularly streaked rufous with dark brown on the edge of feathers (lacks the white outlined by dark brown along the shaft of the adult nonbreeding plumage), back light rufous with dark brown bars, wing coverts and flight feathers light rufous with dark brown bars, rump and upper tail-coverts dark brown with narrow buff bars, tail dark brown with narrow rufous-buff bars (dark bars 6 mm wide, buff bars 3 mm wide); underparts, cheeks rufous, chin and throat whitish, breast whitish buff with fine paler shaft streaks and barred black on the flanks, belly whitish, lower belly gray and under tail coverts blackish with buff bars, under wing coverts pale chestnut with fine black bars; iris brown, bill pale horn, sometimes with a dark culmen. The iris develops a tan outer ring and darker brown inner ring during the time of postjuvenile molt.

NESTLING: Nearly naked at hatching, with thick, stiff, buffy hair-like down; iris gray, mouth lining dark red-pink, changing to pale flesh color with top of tongue black, bill flesh color to brown, feet grayviolet (Swinhoe 1863). Swinhoe identified the coucals on Taiwan, on which this description was based as C. viridis, but C. bengalensis is the only coucal on Taiwan. The yellow-white hair-like down remains attached to the upper back feathers until fledging. SOURCES: AMNH, BMNH, CM, CMNH, FMNH, MCZ, MVZ, MZB, RMNH, ROM, UMMZ, ZMUC, ZRC.

Subspecies Centropus bengalensis bengalensis (Gmelin, 1788), India, Nepal, Bangladesh, Burma south to Tenasserim,Thailand, southeast Asia; Centropus bengalensis lignator Swinhoe, 1861; larger than bengalensis; SE China, Hainan and Taiwan; Centropus bengalensis javanensis (Dumont, 1818); smaller than bengalensis; Malay Peninsula, Greater Sundas (Sumatra, Riau Archipelago, Lingga Archipelago, Bangka, Billiton; Java and Borneo) and the Philippines (Luzon, Mindanao, Mindoro, Cebu, Negros, Balabac, Palawan); Centropus bengalensis sarasinorum Stresemann, 1912; larger and darker than bengalensis; Sulawesi, Banggai Islands, Sula Islands, Lesser Sundas,Timor; Centropus bengalensis medius Bonaparte, 1850; larger than sarasinorum; the Moluccas (except SE), Tanimbar Islands (this form?). Philippine birds have sometimes been recognized as a subspecies C. b. philippinensis Mees 1971, said to differ from C. b. javanensis in the paler margin of the upper wing coverts, and it does in the type series of philippinensis in RMNH, and the form philippinensis was recognized in Dickinson et al. (1991). However, individual variation in adults is as great as betweenpopulation variation and may be related to wear or to age. C. b. philippinensis here is considered a synonym of C. b. javanensis. Birds in the central plains of Thailand, described as C. b. chamnongi Deignan 1955, are intermediate in size between C. b. bengalensis and lignator. Centropus bengalensis has been considered conspecific with African Black Coucal C. grillii, with

252 Lesser Coucal Centropus bengalensis which it is more similar in calls, size and habitat than either coucal is to Madagascar Coucal C. toulou (Dowsett and Dowsett-Lemaire 1993). Molecular genetics indicate that C. bengalensis and C. grillii share a more recent common ancestor than either does with C. toulou, while C. bengalensis is most closely related to Philippine Coucal C. viridis, a species that lives in the same range in the Philippines, than to C. toulou. These coucals each have distinct songs and they are distinct species.

Measurements and weights Centropus b. bengalensis; Assam: Wing, M (n ⫽ 7) 138–150 (145.0 ⫾ 6.0), F (n ⫽ 12) 158–174 (165.9 ⫾ 4.9); tail, M 161–188 (172.3 ⫾ 10.3), F 180–215 (194.0 ⫾ 11.9); bill, M 21.2–24.4 (22.8 ⫾ 1.2), F 22.6–26 (24.4 ⫾ 1.2); tarsus, M 34–38 (36.1 ⫾ 1.9), F 40–45 (42.7 ⫾ 2.2); hallux claw, M 21.2–24.4 (22.8 ⫾ 1.2), F 23.0–25.6 (24.6 ⫾ 1.2) (UMMZ); C. b. lignator; Taiwan, Amoy and Hong Kong: Wing, M (n ⫽ 5) 148–159 (153.2 ⫾ 4.8); F (n ⫽ 11) 165–174 (169.5 ⫾ 3.0) (Stresemann 1912); C. b. javanensis; Salanga, Sumatra, Bali, Natuna, Java, Malacca, the Philippines: Wing, M (n ⫽ 29) 125–147 (133.6 ⫾ 19.0); F (n ⫽ 25) 150–166 (158.9 ⫾ 4.7) (Stresemann 1912); C. b. sarasinorum; Sulawesi, Lesser Sundas,Timor: Wing, M (n ⫽ 5) 146–159 (153.4 ⫾ 5.3), F (n ⫽ 5) 175–188 (178.2 ⫾ 6.0) (RMNH); M (n ⫽ 41) 144–168 (152.9 ⫾ 23.5); F (n ⫽ 31) 169–190 (178.0 ⫾ 5.0); (Stresemann 1912); C. b. medius; the Moluccas: Wing, M (n ⫽ 9) 160–177 (171.7 ⫾ 5.4), F (n ⫽ 12) 190–205 (199.0 ⫾ 3.9) (Stresemann 1912). Weight, C. b. lignator;Taiwan: some mis-sexed?, M (n ⫽ 10) 95–187 (130.3 ⫾ 40.8), F (n ⫽ 10) 138–200 (176.8 ⫾ 17.4) (RMNH); C. b. javanensis, Philippines: M (n ⫽ 4) 82.7–90.0 (86.2), F (n ⫽ 5) 132–160.6 (148.3) (FMNH), C. b. sarasinorum, M (n ⫽ 5) 132–156 (145.6), F (n ⫽ 5) 227–290 (264.4) (ZMB). Wing formula, P5 ⬎ 6 ⬎ 7 ⬎ 4 ⫽ 8 ⬎ 3 ⬎ 9 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length, male 31 cm, female 34 cm (India).A small coucal, in breeding plumage glossy black with

chestnut wings and a long black tail sometimes tipped white; the under wing coverts are pale not black. The bird is much smaller than Greater Coucal Centropus sinensis, which lives in much of its range; C. bengalensis looks streaked and scruffy due to the glossy shaffs of the head feathers, the chestnut color is paler, and the birds tend to be in grassy habitats, not forests. In the Philippines where it occurs with Philippine Coucal C. viridis, C. bengalensis is smaller and the wings are less rufous. In India the nonbreeding birds differ from Sirkeer Malkoha Taccocua leschenaultii in the plumage barring and the bill being brownish not red. In winter (first-year) nonbreeding plumage, above and below pale yellowish brown barred and streaked with dark brown extending onto the tail ( juvenile C. sinensis is darker and barred on the wings).

Voice Calls, (1) series of paired notes,“whoot, whoot”, then a series of “kurook, kurook, kurook” increasing in tempo and descending in pitch. The “whoot” notes are about 0.1 sec, shorter than notes of Greater Coucal C. sinensis. Calls are similar in India, southeast Asia and Malay Peninsula; (2) a series of short notes, “tu-dut” or “tu-dut-dut”, some pairs of notes low (0.4 kHz) and some pairs of notes high (1.0 kHz), in alternation, apparently given by mates (Scharringa 1999). Sometimes the members of a pair call together, each at its own rate, one bird giving a “whoot” and the other a “tu-dut-dut”.Variations on these themes are a series of hollow “booh-booh . . .” breaking into a tinkling cadence, a slowly accelerating series of short high notes “pok, pok, pok, po, po-popo-po-po . . .” descending in pitch, a rapid staccato call like a knock,“totok, totok, totopuk, totopuk . . .” and a coarse, harsh “krah, krah” (Meyer 1879, Coomans de Ruiter and Maurenbrecher 1948, Coomans de Ruiter 1950,Watling 1983, MacKinnon and Phillipps 1993, Smith 1993a,b, Coates and Bishop 1997). Calls are similar across the species range (Smythies 1960, Ali and Ripley 1969, Scharringa 1999, Wells 1999, Kennedy et al. 2000, Robson 2000a, Sheldon et al. 2001, Supari 2003). The common name “Dudut” in Indonesia is onomatopoeic (Bernstein 1859).The breeding adult approaches the nest with rapid short notes (Loke 1953).

Lesser Coucal Centropus bengalensis 253

Range and status Indian subcontinent (lower hills at base of the Himalayas and N Gangetic Plain from northern Uttar Pradesh east to Arunachal Pradesh, NE India in S Assam hills and the Sundarbans and in Bangladesh, and in peninsular India in the Western Ghats in Karnataka, Kerala and Tamil Nadu), southern China, Hainan, mainland southeast Asia, Malay Peninsula, Sumatra (also Riau Archipelago, Lingga Archipelago, Bangka, Belitung), Borneo, Lesser Sundas, Timor, Sulawesi, Moluccas, Tanimbar and the Philippines (Hume and Davison 1878, Rensch 1931, Smythies 1940, 1981, Hoogerwerf 1964, Mees 1971, Eck 1976, White and Bruce 1986, van Marle and Voous 1988, Cheng 1991, Dickinson et al. 1991, Lekagul and Round 1991, Coates and Bishop 1997, Grimmett et al. 1999, Wells 1999, Thomas and Foote 2003). In Sri Lanka the presence of this coucal is questionable; no records are mentioned by Legge (1880) or Baker (1927), and the record in Wait (1925) was based on a purchased skin (Harrison 1999, Thomas and Foote 2003). Resident in most of its range, migratory in parts of China, moving in winter onto the plains of Burma, Thailand and Vietnam (Smythies 1940, Deignan 1955, Robson 2000a), and one was mist netted at floodlights at Fraser’s Hill, in migration at night (Wells 1999).They colonize new sites where forests are cleared (Holmes and Burton 1987). After the 1883 volcanic eruption and devastation of all plant and animal life on Krakatau, these coucals had colonized the re-vegetated island by 1908 from a source at least 15 km distant over open water, and

they nested on the island soon afterwards (Dammerman 1922, Robinson and Kloss 1923b, Thornton 1996). Common in much of the range. Longest known survival from date of ringing, 74 months (McClure 1998).

Habitat and general habits Tall wet grass including elephant grass Imperata, reedbeds, standing rice fields, grasslands, agricultural plains, sugar cane fields, swamp, bamboo thickets, second growth forest, open country scrub and cultivation; mainly in lowland floodplains and lower valleys. They get into submontane tracts of the Himalayas, and they are widely distributed in the Indian Peninsula and in Bangladesh.They live from the lowlands to 365 m, and occasionally to 1400 m in Nepal, in isolated hill-station agriculture to 1500 m in the Malay Peninsula, to 1500 m in Sabah (most records below 700 m, once to 1500 m), 700 m in Sangihe, 1000 m in Sulawesi, 820 m in Buru, 650 ⫹ m in Seram, 900 m in Lombok, 900 m in Sumbawa, 1500 m in Flores and 500⫹ m in Sumba, rarely higher ( Jerdon 1862, Meyer 1879, Inskipp and Inskipp 1985, Coates and Bishop 1997, Riley 1997, Bishop and Brickle 1998, Wells 1999, Kennedy et al. 2000, Sheldon et al. 2001).Terrestrial, they do not climb about in trees (Smythies 1981). Heard often, they are seen in the early morning and late afternoon when the coucals are in the open on sunning perches, look-outs and song posts, or in labored flight low over the vegetation.

Food Insects, including grasshoppers (Acrididae), locusts, crickets, mantids, beetles, hemiptera, hairy caterpillars, Sphingid caterpillars; also spiders, lizards (Calotes), frogs and fruit (Ali and Whistler 1937, Loke 1953, Rabor 1977, Smythies 1981, Sody 1989,Wells 1999; ZMUC).

Displays and breeding behavior Monogamous, live in pairs. In courtship feeding, the male holds a large insect in his bill.The female approaches him, raises her head, neck and bill, and calls “tok tok tok tok” and the male approaches her, raises his tail and bows to her. Copulation lasts about 30 sec. In India, both sexes are said to build

254 Violaceous Coucal Centropus violaceus the nest, incubate and care for the young (Baker 1934, Ali 1953, 1996), whereas in Java, the male alone is said to build the nest, incubate and feed the young (Bernstein 1859, Spennemann 1928).

Breeding In India they breed from May to September, in the plains they nest after the rains begin in June, in the hills they begin by April and May, in Kerala they breed in August and September (Hume 1873, Baker 1934, Ali 1953); in Burma they nest in the rains (Oates 1877, Smythies 1981), in Thailand they breed from May to August (Riley 1938), in the Malay Peninsula eggs are laid from December to July and fledglings appear into October (Wells 1999), in Taiwan they nest in September (Swinhoe 1863), in the Philippines from March to May (Rabor 1977), in Java from January to July, and in November (Spennemann 1928, Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967), in Flores they

breed from February through September, mainly in March and April (Verheijen 1964), in Sarawak at the time of the rice harvest in February and in Sabah from December to May (Smythies 1999, Sheldon et al. 2001). The nest is built near the ground in dense vegetation, grass or low bushes, well concealed, c. 25 cm h i g h ⫻ 18 cm across. The round, ball-like nest is formed with live grasses pulled down and shaped into a nest which is then covered with a dome of grass, often lined with green leaves, and has a side entrance (Jerdon 1862, Swinhoe 1863, Hume and Oates 1890, Herbert 1924, Loke 1953, Wells 1999, Sheldon et al. 2001). Eggs are chalky white, nearly round, 28 ⫻ 24 mm (India), 30.3 ⫻ 24.8 mm (Malay Peninsula), 27.6 ⫻ 23.8 mm ( Java), 31 ⫻ 24 (Labuan) and 32 ⫻ 25 mm (Flores). The clutch is usually 3(2–6) in India, 2–4(5–6) in southeast Asia, 2(3) in Malay Peninsula, and 4 in Taiwan (Sharpe 1879, Riley 1938, other references above). Incubation and nestling periods are unknown.

Violaceous Coucal Centropus violaceus Quoy and Gaimard, 1830 Centropus violaceus Quoy and Gaimard, 1830, Observations zoologiques faites à board ‘de l’Astrolabe , 1: 229; Atlas, Oiseaux, pl. 19. (Carteret Harbor, New Ireland) Other common names: Giant Forest Coucal. Monotypic.

papillated ridge along the palate cleft, tongue pink with a papillated white rear edge and scattered white papillae on the dorsal surface, and a black U-shaped band near the tip, feet bluish gray (Gilliard 1961).

Description

Measurements and weights

ADULT: Sexes alike, upperparts and underparts black with violet gloss, wing black with violet gloss, tail long, graduated, broad and black with violet gloss; bare skin around eye white to light reddish, eye-ring black, iris red, bill black, feet pale slaty horn.

Wing, M (n ⫽ 6) 258–292 (271.2 ⫾ 12.5), F (n ⫽ 5) 245–288 (261.2 ⫾ 16.0); tail, M 340–405 (379.6 ⫾ 18.4), F 310–400 (356.0 ⫾ 34.9); bill, M 61.5–69 (65.3 ⫾ 2.9), F 62–77 (69.1 ⫾ 5.9); tarsus, M 61.5– 68 (65.3 ⫾ 2.9), F 62.5–70 (66.3 ⫾ 3.2), hallux claw, M 20.5–25 (22.3 ⫾ 1.7), F 20–23 (21.0 ⫾ 1.41) (AMNH, ZMUC). Weight, F (n ⫽ 1) “500 g” (Gilliard and LeCroy 1967a), label of specimen (AMNH 777857) says “500⫹ g”, perhaps the capacity of the weighing scale. Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⫽ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

JUVENILE: Upperparts dull black, wing and tail black glossed violet; underparts loose-webbed, sooty gray; iris light gray. NESTLING: Black skin, long white feather sheaths to 35 mm on head, back and wing, bill black with a white tip, mouth lining purplish pink with white

SOURCES: AMNH, ZMUC.

Lesser Black Coucal Centropus bernsteini 255

Field characters Overall length 64–70 cm. A very large blackish coucal with a violet gloss, a long graduated tail and a red eye.The bare white skin around the face and the whitish feet show clearly in the field.

Voice Deep hollow booming notes and loud booming duets “oo-OOMP”, given in a continuous series; the calls are loud and carry over several kilometers (Coates 2001, Mayr and Diamond 2001).

Range and status Bismarck Archipelago: New Britain and New Ireland, absent from the smaller islands (Rothschild and Hartert 1907, Hartert 1925c, Coates 1985, Mayr and Diamond 2001). Resident: More common in New Britain than in New Ireland (Coates 2001). Sparsely distributed both in lowlands and in the mountains; it is estimated to be one of the 25 rarest species in the Bismarck Archipelago, with fewer than 1000 pairs in 10,000 square kilometers (Mayr and Diamond 2001).

about in tree branches and vines to a height of 15 m, and they plane from the top of trees to lower levels (Hartert 1926). One bird walks to another adult carrying a large insect in the bill, in courtship feeding (Gilliard and LeCroy 1967a). Occur singly and in pairs; the members of a pair allopreen (Coates 2001).

Habitat and general habits

Breeding

Primary tropical rain forest, disturbed habitat, active in limbs and vines and ferns, from lowlands to 950 m. Call loudly back and forth in forest, with neck pulled down and tail hanging, as they give booming calls.They keep in cover of vegetation and are difficult to observe. They jump, hop and walk

Season is November to January. Nest is a loose structure of twigs built at the top of a tall tree in the forest. Eggs are white, somewhat glossy, 43 ⫻ 34 mm, the clutch is 2 or 3 (Hartert 1926, Meyer 1931, 1933, 1936, Gilliard and LeCroy 1967a). Incubation and nestling periods are unknown.

Food Large insects (stick insect as large as 18 cm), frogs, and tiny snails (Gilliard and LeCroy 1967a).

Lesser Black Coucal Centropus bernsteini Schlegel, 1866 Centropus Bernsteini Schlegel, 1866, Nederlandsch Tijdschrift voor de Dierkunde, 3, 251. (New Guinea ⫽ Vogelkop peninsula) Other common names: Black Scrub Coucal, Bernstein’s Coucal. Monotypic.

Description ADULT: Sexes alike, a small, short-tailed black coucal with black shaft streaks, upperparts glossed green, wing black, tail black, wing all black below, under

wing coverts black; iris dark brown, bill black, legs and feet black.A nonbreeding adult plumage is unknown. JUVENILE: Above blackish narrowly (1–2 mm) barred buff or rufous, wing and tail blackish with narrow (2.5 mm) buff to pale rufous bars, tail feathers narrow (32 mm wide, vs 42 mm in adult), throat whitish in midline, neck and breast marked with chestnut brown on sides, central belly gray with blackish bars, remaining underparts blackish brown with narrow pale barring.

256 Lesser Black Coucal Centropus bernsteini NESTLING: Short white tips on growing feathers of the face (AMNH 267142), not spiked with long hair-like down as in some other coucals. SOURCES: AMNH, ANSP, BMNH, FMNH, MSNG, MVZ, SMTD, ZFMK, ZMB, ZSM.

Subspecies No subspecies are recognized. In one sample (AMNH), the birds on Manam were smaller than birds in New Guinea and were described as a subspecies Centropus bernsteini manam, (Mayr 1937). In BMNH and MSNG samples, males on Manam and New Guinea overlap in size.

Measurements and weights New Guinea: Wing, M (n ⫽ 12) 164–193 (173.8 ⫾ 8.5), F (n ⫽ 8) 176–196 (180.0 ⫾ 8.0); tail, M 220–265 (247.4 ⫾ 13.6), F 230–265 (247.8 ⫾ 10.9); bill, M 27–33 (31.3 ⫾ 6.0), F 28–35 (31.8 ⫾ 6.2); tarsus, M 39–44 (41.2 ⫾ 1.9), F 38–44 (41.4 ⫾ 2.3); hallux claw, M 21–26 (23.9 ⫾ 1.4), F 24–26 (25.1 ⫾ 0.6) (AMNH, BMNH, ZSM); Manam: Wing, M (n ⫽ 4) 179–190 (182.8 ⫾ 5.0), F (n ⫽ 4) 188–200 (191.8 ⫾ 8.4); tail, M 248–275 (261.2 ⫾ 11.1), F 271–277 (274.5 ⫾ 2.5) (Mayr 1937 (males), AMNH, BMNH). Weight, M (n ⫽ 5) 130–160 (146), F (n ⫽ 2) 160–200 (180) (AMNH, FMNH, Hartert 1930). Wing formula, P5 ⫽ 4 ⬎ 3 ⱖ6 ⬎ 7 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 46–52 cm. Small black coucal with a short tail, dark iris and a black bill. This species is smaller and more uniformly black than the Pheasant Coucal C. phasianinus with which it occurs in New Guinea, and has a dark iris (unlike yellow iris in Biak Coucal C. chalybeus).

Voice Three hoots, “woop woop woop” on a descending scale, sometimes given in longer series. Pairs often duet, one bird with “hui” and the other with “whui”, a lower note given at a slower rate. Call is weaker and flatter than the call of Greater Black Coucal C. men-

beki and shorter than that of Pheasant Coucal C. phasianinus in New Guinea (Gyldenstolpe 1955, Coates 1985, Beehler et al. 1986,Tolhurst 1992).

Range and status Western and central New Guinea; Manam I. In New Guinea they occur in the west from Vogelkop, and north from the Idenburg River and Humboldt Bay east along the coast to Sepik River, along the coast of the Huon Peninsula, Lae and Bulolo, and in the south from the Mikima and Setikwa rivers in the west and the Fly and lower Turama rivers near Kikori in the east. Resident. Largely allopatric with Pheasant Coucal C. phasianinus, but both occur in the north. Not uncommon in second growth and cane grass, they are shy and difficult to observe except when they perch in the open to call in the morning and late afternoon (Maxwell 1938, Beehler et al. 1986, Tolhurst 1992, Coates 1985, 2001).

Habitat and general habits Scrub, tall cane grass, rank vegetation and edge of forest along rivers and lakes; sea level to 500 m, rarely to 900 m (Mayr 1941, Rand 1942b, Gilliard and LeCroy 1966, Rand and Gilliard 1967, Coates 1985, Beehler et al. 1986).

Food Small lizards, small snakes, grass insects, butterflies (D’Albertis and Salvadori 1879, Gyldenstolpe 1955, FMNH).

Pheasant Coucal Centropus phasianinus 257

Breeding In New Guinea region the season is prolonged: a nest with eggs was found in early May along the Idenburg River (Rand 1942b), where a young bird with a short tail was found in December (AMNH 207147), a laying female was taken at Kaku in November (FMNH 267143), and a female had a large ovary in August at Mimika River, Wkatimi (BMNH 1911.12.20.961). Along the Sepik River, local people say the birds nest at the time of low water in July (FMNH 280092). Breeding males

sometimes have two large testes (FMNH), in contrast to some other coucals, while other males of this species have a rudimentary left testis (Rand 1942b).The nest is a covered mass of cane grass and has an irregular untidy opening in the side, and is built in grass and supported on all sides by the stems of the grass. One nest measured 21 ⫻ 38 cm high outside, 15 ⫻ 25 cm inside. Eggs are white, 32 ⫻ 25.5 mm, the clutch is 2 (Rand 1942b). The nest is attended to by the male (AMNH 339693). Incubation and nestling periods are unknown.

Pheasant Coucal Centropus phasianinus (Latham, 1801) Cuculus phasianinus Latham, 1801, Index Ornithologicus, Suppl.: xxx. (New Holland [⫽ New South Wales]) Polytypic. Six subspecies. Centropus phasianinus phasianinus (Latham, 1801); Centropus phasianinus spilopterus G. R. Gray, 1858; Centropus phasianinus nigricans (Salvadori, 1876); Centropus phasianinus thierfelderi Stresemann, 1927; Centropus phasianinus propinquus Mayr, 1937; Centropus phasianinus mui Mason et al. 1984. Other names: Australian Pheasant Coucal, Swamp Pheasant, Kai Coucal, New Guinea Pheasant Coucal, Common Coucal (New Guinea).

Description ADULT: (C. p. phasianinus): Sexes similar, breeding plumage, the head, upper back and underparts dull black, feather shafts glossy blacks, wing coverts rufous brown with black and cream barring and bold white streaks, wing flight feathers with inner webs rufous, outer webs barred buff, black and rufous brown, showing a broad dark trailing edge in flight, back, rump and upper tail coverts blackish brown with fine pale barring, tail long, black with buff bars, the broad bars sometimes tricolor with rufous bars inserted into the black bars, and the rufous bars sometimes spotted with black, tail feathers broad and rounded with a white tip; under wing coverts black (greater coverts) to rufous brown with fine blackish bars (lesser coverts).

Eye-ring narrow and blackish, iris red (breeding males) to pale brown (nonbreeding males), orange to yellow (breeding females), or whitish (nonbreeding females), bill black (breeding) to pale horn (nonbreeding season), feet black. Nonbreeding plumage, the upperparts rufous brown with buff streaks, tail brown with gray bars and feathers pointed at the tip, tail longer than in breeding plumage, underparts buff with paler streaks. In New Guinea and Kai Island subspecies, upperparts and underparts dull black, wing coverts and flight feathers sometimes with buff spots and bars, long tail black glossed green, under wing coverts black; iris red, bill black, feet black. Nonbreeding plumage, perhaps restricted to first year, upperparts dull black with buff streaks on head and upper back, wing blackish with narrow buff bars, tail black with narrow buff bars, chin whitish, throat and breast to belly whitish buff with straw-colored rachi, outer breast with fine black spots and bars, flanks black, lower belly and under tail coverts blackish with buff bars, bill pale yellowish (AMNH 628289). Birds molt from juvenile plumage into this plumage, and they molt into the black adult plumage. JUVENILE: Plumage black and rufous, head nearly all black (not streaked), back black with whitish buff spots or bars, wing rufous with narrow black bars, outer vane of primaries buff with black and rufous bars, inner vane rufous with whitish bars, tail

258 Pheasant Coucal Centropus phasianinus blackish with rufous bars, chin and center of throat whitish with dark shaft streaks, throat to belly blackish with whitish spots on throat and whitish mottled bars on breast and belly; iris brown, bill brown to pale below. Regional variation in plumage and size is less than in the adults in New Guinea and Kai Islands, plumage blackish with whitish buff spots and bars, feather shafts on head and upper back glossy black, tail black with narrow buff bars, chin white, underparts blackish with white bars, overall appearance black with whitish spots; iris gray to brown. NESTLING: Skin black with pink abdomen, upperparts with white hair-like natal down, most conspicuous on top of head and neck where the bristles are directed forward; palate red with pale patch of papillae on either side of the palate cleft, tongue red with a white papillated rear edge and a black U-shaped band near the tip. SOURCES: AMNH, BMNH, CM, FMNH, MSNG, MSNM, MZB, RMNH, ROM, SMTD, UMMZ,WAM, ZMB.

History and subspecies Centropus spilopterus was formerly considered a species restricted to the Kai Islands. Salvadori (1876, 1881) described the form nigricans of New Guinea as a distinct species as it had pale marks on the wing. New Guinea coucals were known as C. spilopterus beginning with Sharpe (1878); Salvadori recognized them as distinct species, C. nigricans. Later works considered the New Guinea forms (nigricans and others) conspecific with Australian Pheasant Coucal C. phasianinus but did not discuss the Kai Is birds (Stresemann 1927b, Mayr 1937, 1941, Mayr and Rand 1937, Mason et al. 1984). Most Kai Island and New Guinea birds are nearly identical in adult plumage pattern, black with pale wing marks (the marks are dark rufous in thierfelderi). White and Bruce (1986) likewise noted that some Kai birds have buffy barring on the flight feathers. Feathered nestlings and juveniles in the Kai Islands (AMNH, MSNG) have a whitish throat and are barred with buff and tan, like the young in New Guinea and Australia. Both spilopterus and nigricans are nearly

identical in mitochondrial gene sequence to C. phasianinus of Australia (Figure 5.3). C. p. spilopterus G. R. Gray, 1858; large size, tail unbarred black, primaries unbarred black, sometimes with pale bars, these narrower than the black bars on the wing, underwing coverts black; Kai Islands; C. p. nigricans (Salvadori, 1876) (includes C. s. obscuratus Mayr 1937); smaller than spilopterus, tail black with narrow indistinct buff bars, primaries black with narrow buff bars (buff bars 1–2 mm, black bars 7–9 mm); southeastern New Guinea, Yule I, D’Entrecasteaux Islands; C. p. propinquus Mayr, 1937; smaller than nigricans, plumage similar to nigricans; northern New Guinea; C. p. thierfelderi Stresemann, 1927; tail black or black with narrow rufous bars, primaries with rufous bars nearly as wide as black bars (4–6 mm, vs 5–8 mm), most birds have rufous to buff bars on the primary coverts, under wing coverts black with buff edge; southern New Guinea. C. p. phasianinus (Latham, 1801); tail black with broad rufous bars, primaries with rufous bars wider than the black bars, breeding plumage with breast black, nonbreeding plumage streaked brown and buff; Australia; C. p. mui Mason et al. 1984; breast white; East Timor (Los Palos district). In Australia, northern and northwestern birds are often recognized as a distinct subspecies, C. p. melanurus (Mason et al. 1984, Mason 1997). Goodwin (1974) and Storr (1977) questioned whether more than one subspecies occur in Australia; Mason et al. (1984) and D. Rogers (in Higgins 1999) found barring on the rectrices to differ between western and eastern birds. Birds in northeastern Australia intergrade with birds south of Ayr, Queensland, and intermediates occur in central coastal Queensland from 20° to 23°S (Mason et al. 1984, Ford 1987a,b, Higgins 1999). The form melanurus is not recognized, because northern and northwestern populations overlap those in eastern and southern Australia in size and in the pattern on the tail, which tends to have wider black bars in the north and northwest (AMNH). Australian birds have a seasonal change of plumage in the adult; birds in

Pheasant Coucal Centropus phasianinus 259 New Guinea and the Kai Islands are not known to have this, although they have a streaked nonbreeding plumage which is worn between the juvenile and first breeding adult plumage; these birds have been considered a distinct species, the New Guinea Pheasant Coucal C. spilopterus or C. nigricans. The Timor form C. p. mui is known only from the type specimen. Although White and Bruce (1986) suggest that mui is a distinct species, variation in the extent of white in plumage within other species of coucals, some with white color morphs (C. goliath, C. ateralbus, C. viridis) is consistent with the white breast in C. p. mui representing a variant within C. phasianinus. The nonbreeding adult plumage is well known in C. phasianinus in Australia; it has not previously been recognized in coucals in New Guinea and the Kai Islands (Mason et al. 1984). Juvenile spilopterus, nigricans and thierfelderi have black plumage marked below with spots and bars of white and buff, and head and upper back black or dark brown with dark feather shafts; the feathers of the crown are sometimes broadly spotted with buff. Nonbreeding adult plumage in these forms has straw-colored streaks on the head and upper back, the streaks being the shaft and base of the barbs; the underparts are pale with straw-colored shaft streaks on the breast, as in nonbreeding adult C. p. phasianinus in Australia. For C. p. thierfelderi, AMNH 425934 is a specimen in molt from mottled black plumage to straw-colored hackles; AMNH 425934 and 421865 are specimens in molt from straw-colored hackles to breeding plumage. For C. p. nigricans, AMNH 781441 is a worn juvenile in molt to non-breeding adult plumage, and AMNH 628289 and MSNG 13975 are birds in straw-hackled non-breeding adult plumage; breeding plumage adults include MSNG 13972, 213973, 13974 and 13976 and AMNH 628288, all in Salvadori’s type series. For C. p. spilopterus, fledged juveniles include AMNH 628279 and MSNG 13969, nonbreeding adults include AMNH 628276, and most birds in AMNH, MSNG, RMNH, SMTD, ZMA are in adult breeding plumage.A black spilopterus with pale marks on the wing, MSNG “d” ⫽ 13968, that Salvadori (1881) thought was a juvenile (‘giovani’), is adult.

Measurements and weights C. p. spilopterus: Wing, M (n ⫽ 9) 212–241 (228.9 ⫾ 11.2), F (n ⫽ 4) 249–268 (256.0); tail, M 271–330 (296.6 ⫾ 18.2), F 290–334 (324.3); bill, M 36–42 (38.5 ⫾ 1.8), F 36–42.5 (39.0); tarsus, M 44–52 (49.6 ⫾ 2.3), F 48–53 (50.8); hallux claw, M 23–27 (25.0), F 21.5–28 (27.3) (AMNH, BMNH, MSNG, RMNH, SMTD, ZMA); C. p. nigricans: Wing, M (n ⫽ 10) 200–229 (216.8 ⫾ 9.8), F (n ⫽ 10) 216, 220, 230–248 (236.3 ⫾ 11.5, including the two small birds, perhaps mis-sexed); tail, M 288–326 (299.0 ⫾ 22.4), F 280, 318, 326–354 (332.0 ⫾ 21.9, including the two small birds, perhaps mis-sexed) (AMNH, MSNG); C. p. propinquus: Wing, M (n ⫽ 5) 194–219 (202.4 ⫾ 10.1) (excludes “M” 248), F (n ⫽ 5) 207–237 (228.8 ⫾ 12.9) (excludes “F” 197); tail, M 260–301 (280.6 ⫾ 18.5), F 287–332 (308.6 ⫾ 17.8) (AMNH); C. p. thierfelderi: Wing, M (n ⫽ 10) 205–222 (214.3 ⫾ 5.8), F (n ⫽ 8) 222–251 (238.3 ⫾ 10.6) (excludes “F” 218); tail, M 264–338 (289.9 ⫾ 21.6), F 282–326 (317.3 ⫾ 19.2) (excludes “F” 278) (AMNH); C. p. phasianinus, Cape York: Wing, M (n ⫽ 9) 217–246 (237.3 ⫾ 9.6), F (n ⫽ 15) 247–274 (260.0 ⫾ 8.7); tail, M 330–378 (357.1 ⫾ 17.9), F 326–418 (376.1 ⫾ 29.3); hallux claw, M 20–27 (23.3 ⫾ 2.6), F 20–26 (23.2 ⫾ 1.4) (AMNH); NSW to central Queensland south of 21°S: Wing, M (n ⫽ 23) 205–248 (225.6 ⫾ 10.3), F (n ⫽ 31) 235–267 (249.3 ⫾ 8.3); tail, M (n ⫽ 13, breeding plumage) 282–342 (310.2 ⫾ 18.7), (n ⫽ 6, nonbreeding plumage) 323–356 (338.3 ⫾ 12.3), F (n ⫽ 12, breeding plumage) 305–361 (337.1 ⫾ 21.2), F (n ⫽ 9, nonbreeding plumage) 333–393 (360.2 ⫾ 22.1); bill, M 33.1–37.2 (35.5 ⫾ 1.3), F 35.2–42.6 (38.6 ⫾ 1.7); tarsus, M 46.3–53.8 (50.5 ⫾ 1.6), F 49.5–56.7 (53.2 ⫾ 1.8) (Higgins 1999); Western Australia: Wing, M (n ⫽ 10) 230–264 (245.2 ⫾ 10.2), F (n ⫽ 7) 260–283 (266.6 ⫾ 8.9); tail, M 348–376 (360.6 ⫾ 16.2), F 360–445 (384.4 ⫾ 26.5) (WAM); Pilbara to Cape York and tropical Queensland: Wing, M (n ⫽ 41) 227–254 (240.7 ⫾ 7.2), F (n ⫽ 40) 241–303 (268.8 ⫾ 12.9) (Higgins 1999);

260 Pheasant Coucal Centropus phasianinus C. p. mui: Wing, M (n ⫽ 1) 250; tail, M 377 (in partial molt) (Mason et al. 1984). In C. p. phasianinus, tail length differs between breeding and nonbreeding adults, longer in the nonbreeding plumage, due to a double annual molt of T1; the other rectrices molt only once a year (Mason et al. 1994, Higgins 1999). Weight, C. p. propinquus, M (n ⫽ 3) 180–225 (205), F (n ⫽ 1), 300 (Hartert 1930); C. p thierfelderi, M (n ⫽ 2) 250–281 (265.5), F (n ⫽ 1) 375 (Mees 1982); C. p. phasianinus: NSW to Central Queensland, M (n ⫽ 16) 200–364 (302.1), F (n ⫽ 22) 242–520 (444.9); Pilbara to Cape York, M (n ⫽ 13) 289–378 (327.6), F (n ⫽ 12) 390–600 (483.0) (Higgins 1999). Wing formula, C. p. spilopterus, P6 ⫽ 5 ⫽ 4 ⬎ 7 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10; C. p. phasianinus, P5 ⫽ 6 ⬎ 7 ⫽ 4 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 1.

Field characters Overall length 60–80 cm. Large coucal, a skulking long-tailed pheasant-like bird, in Australia rufous above and the head, neck and underparts glossy black in breeding season, streaked above and below in nonbreeding plumage, long barred tail; in New Guinea and the Kai Islands, dull black plumage, often with pale spots on wing coverts and flight feathers, and a long, glossy green tail.

Voice Dull rapid resonant booming series of “hooh” notes, falling then rising in pitch, “coo-coo-coocoo-coo-coo-coo-coocal”, accelerates in tempo, with 20 or more notes in a series and lasting up to 6 sec. Members of a pair often sing in duet. Also gives a softer series of “coo” notes all on same pitch and lasting 3–4 sec. Other calls include harsh scolding “keouw”, a dull thud “pthuk”, a double note “nah-oo” used when danger appears, a grunt when moving through undergrowth and approaching a nest, and a hiss when alarmed. Begging calls of the young sound like shimmering, buzzing, or a loud clicking; and nestlings give a snake-like hiss “pssch” (Slater 1971, Mackness 1979, Buckingham and Jackson 1990, Taplin and Beurteaux 1992, Higgins 1999). Calls in New Guinea are like calls in

Australia (Diamond 1972, Coates 1985, 2001, Beehler et al. 1986, Coates and Bishop 1997).

Range and status New Guinea, Kai Islands, Australia and Timor. Resident. Conspicuous in breeding season. Kai Islands (Kai Kecil, Kai Besar, Roa). In New Guinea, local in the west in Irian Jaya (in the south at Merauke and Pulau Kinaan, in the north from Mamberano River to Jayapura at Humboldt Bay), in the east widespread in the lowlands; D’Entrecasteaux Islands (Goodenough, Fergusson and Normanby), Yule I and nearshore islands in the Torres Strait (Boigu, Daan, Saibai) (Hartert 1930, Mayr 1941, Mees 1982, Draffan et al. 1983, Beehler et al. 1986, White and Bruce 1986, Sibley and Monroe 1990, Coates and Bishop 1997, Mason 1997, Coates 2001). Common in lowlands. The Kai Islands form has been considered near-threatened (Collar et al. 1994). Not reported on the remote islands in the Torres Strait, and there is no evidence of dispersal between New Guinea and Australia. In Australia, common near the coast, uncommon in semiarid areas and scarce in the arid interior (Blakers et al. 1984, Ford 1987b, Mills 1987, Higgins 1999); in Western Australia occurs south to Pilbara Region (Storr 1984a); the last records in the Gascoyne Region were in the late 1800s (Storr 1985). In recent years, coucals have decreased in numbers in Australia (Seventy and Whittell 1976, Johnstone and Storr 1998, Higgins 1999). Population density, in southern New Guinea woodlands, c. 0.1 bird /ha (Bell

Pheasant Coucal Centropus phasianinus 261 1982a); in Australia, 0.02–0.4 birds / ha, highest in Kakadu NP (Higgins 1999).

Rand 1937, Cooper 1953, Frauca 1974, Coates 1985, Higgins 1999).

Habitat and general habits

Displays and breeding behavior

In New Guinea, coucals occur in the forest edge, lightly wooded cultivation, savanna, remnant forest patches and scrub, common in cane and grass, also in reed-bed; found in lowlands below 500 m, and locally in open highland valleys as high as 1500 m and 1800 m (Diamond 1972, Mason et al. 1984, Beehler et al. 1986, Coates 2001). In Australia, they live in dense riverine vegetation, long grass, rank herbage, coastal heathlands, margins of swamps, canefields, thickets of Melaleuca, Lantana and Pandanus, thickets of vines, mangroves, secondary forests, spinifex in sandstone country, sandstone gorges, dense grass along roads, and gardens. In Timor, they are in swampy grassland, monsoon forest and fringing grasslands (Mason et al. 1984). The coucals live singly and in pairs. They are mainly terrestrial, and run on the ground, with body low, neck stretched and head low and held forward. They climb into trees and perch, where they call from horizontal branches in midstory and lower crown at forest edge. They forage by clambering about on the ground and through thickets in search and pursuit of large insects. When disturbed they run or fly into cover (Bell 1970, Mackness 1979, Coates 1985, Coates and Bishop 1997, Higgins 1999). They search for food as they walk in low dense vegetation and ground cover, then run down the prey in “flush and rush” behavior, tail extended and body held low. Like other cuckoos, they are conspicuous when they sunbathe, spreading the back feathers to expose the skin, and spreading the wings and tail to dry after a rain or after foraging in wet vegetation. When feeding on large prey, such as a large snail, the bird holds the prey under its foot and tears it apart with its bill.

Socially monogamous, adults live in pairs, territorial in the breeding season. Coucals perch in the open when they call in the morning and late afternoon. Both members of a pair call in duet (Taplin and Beurteaux 1992, Higgins 1999). The male sometimes carries food and feeds the female as they mate. In courtship feeding, one bird holds a food item in its bill while it raises and spreads its tail, then moves onto the partner and mates, feeds the mate and then flies into tall grass. In copulations when no courtship feeding is observed, the male crouches behind the female, moves his head from side to side or up and down as he follows her; she runs ahead, keeping her body low, then the male straightens his body and drags his wings along the ground and spreads his tail, the female raises her body and the male mounts. Courtship feeding has been seen in New Guinea and Australia (Crawford 1972, Mackness 1979, Coates 1985, Higgins 1999). In Australia, both female and male incubate the clutch, and both feed the young; the male does most feeds (87% of 151 feeding visits). Both parents remove fecal sacs from the nest (Taplin and Beurteaux 1992).

Food Large insects (grasshoppers, crickets, mantids, stick insects, caterpillars, beetles, bugs), Sesarma mangrove mud crabs, spiders, scorpions, snails; frogs, lizards, snakes, nestling birds, small mammals including rodents and bandicoots (Mathews 1918, Mayr and

Breeding In the Kai Islands a nestling was taken in July (Hartert 1903) and fledglings with 12-mm short tails were taken in April and September (AMNH). In southern New Guinea (Fly River to Port Moresby lowlands and the Sogeri Plateau), adults carry nest material in May (Coates 1985), a nest had young in June (Bell 1969), nests have eggs in September (Coates 1985), tail-less fledged young appear in early November (Coates 1985) and a female had an egg in the oviduct in late December (AMNH 425929), all indicating a long period of nesting in the dry season. In Australia they breed near Sydney from October to December (Cooper 1953), in Queensland from September to May (Storr 1984b), in Kimberley Division from November to March (Storr 1980), in Northern Territory from December to April (Storr 1977).

262 Sumatran Ground-cuckoo Carpococcyx viridis The nest is a covered mass of grass, open at either end, like a shopping basket with a broad handle over the top. It is begun as an open platform, then the blades and stems are drawn together to form a cover, then a nest lining is added below with sticks as long as 25 cm, then fibers and leaves. It sometimes has an entrance tunnel or a platform of leaves near the entrance.The nest is built in a grass tussock near the ground or in a bush or pandanus. Eggs are white, becoming brown with stain; in New Guinea 36 ⫻ 27 mm, in Australia 38 ⫻ 29 mm; clutch 2–7, usually 3–5 (Hartert 1903, Hindwood 1942, 1957, Cooper 1953, Schönwetter 1964, Frauca 1967, Pratt 1972, Beruldsen 1980, Storr 1984a,b, Taplin and Beurteaux 1992). Adults incubate with the first egg, and they add green leafy twigs through incubation and after young have hatched. Incubation period is 15 days. Nestlings hatch asynchronously, a day apart, with long white

hair-like down on the top of head, also on back; skin black, gape bright red, edge of tongue black. Nestling pinfeathers break through the skin at 6 days when the pins are 2.5 mm long on the wing, at 11 days, feather brushes emerge from pins; the eyes open at 7 days (Frauca 1967, Pratt 1972, Hoppe 1988,Taplin and Beurteaux 1992). In New Guinea a short-tailed young was carried by the parent in flight as the adult held the body of the young in its claws (Bell 1969, 1970, 1984). Nestlings weigh 18.6 g at hatching and 123.5 g at fledging (male?), or 41% of adult male weight. Undisturbed nestlings fledge in 17 to 21 days (Cooper 1953). The natal down is gone within a week of fledging, and the young are independent at 40 days.A female may lay several clutches and rear three to four broods in a season (Mackness 1979). Breeding success, of 14 nestlings, 11 (79%) fledged, others were taken by a predator (Frauca 1967,Taplin and Beurteaux 1992).

Couinae Genus Carpococcyx G. R. Gray, 1840 Carpococcyx G. R. Gray, 1840. A List of the Genera of Birds, p. 56. Type, by monotypy, Calobates radiceus Temminck 1832.The genus name refers to the fruiteating behavior (Gr. karpos, fruit; kokkux, the cuckoo)

and the birds are also known as fruit-cuckoos. The genus is characterized by large body size, bare skin on the face, and gray plumage with iridescent gloss; the group is restricted to the Asian tropics.Three species.

Sumatran Ground-cuckoo Carpococcyx viridis Salvadori, 1879 Carpococcyx viridis Salvadori, 1879, Annali del Museo Civico di Storia Naturale di Genova, 14, p. 187. (Sumatra) Other common names: Sunda Ground-cuckoo (in part), Green-billed Ground-cuckoo (in part). Monotypic.

Description ADULT: Sexes alike, crown black shading to green on hind crown, upper back, wing coverts and secondaries dull green, lower back and rump dark rufous with fine black bars, wing and tail glossy

greenish-black; underparts, black patch from chin to under the eye extends to side of crown behind the bare facial skin, throat and upper breast dull greenish gray, lower breast and belly rufous buff finely barred with black, flanks and under tail coverts rufous barred black, under wing coverts brown; large area of bare skin on face, above the eye pale green, behind eye lilac and above the cheek crimson red, iris red, bill blackish above and pale green below, tarsi green. JUVENILE: Different from adult, upperparts dark rufous with indistinct black bars, rump paler rufous

Sumatran Ground-cuckoo Carpococcyx viridis 263 with black bars, wing coverts barred rufous and dark brown, alula black, feathers with a broad buffy brown edge, wing dark brown with feathers edged rufous brown, tail dark brown; underparts with the throat and breast rufous indistinctly barred with black, belly rufous indistinctly barred whitish and black, under tail coverts rufous barred black; bare skin around eye less extensive than in adult, iris dark gray, bill black, feet dark gray. NESTLING: Undescribed. SOURCES: MSNG, RMNH, ZRCNUS. The bird was first collected in Sumatra, a specimen that Schlegel identified as the previously unknown female of the Bornean Ground-cuckoo C. radiatus. When Salvadori obtained specimens taken by Odoardo Beccari on Gunung Singgalang in the Padang highlands, he described the bird as distinct from C. radiatus (Salvadori 1879, Finsch 1898). Sumatran Ground-cuckoos are known from only nine museum specimens (MSNG, RMNH, SMNS, ZRC; Arbocco et al. 1978, Collar and Long 1996, BirdLife International 2001) and had not been seen by western ornithologists since 1916, the last known specimens, when they were rediscovered in a forest in 1997 (Zetra et al. 2002).

Measurements Wing, M (n ⫽ 2) 210 (210), F (n ⫽ 1) 210, U (n ⫽ 2) 206, 211; tail, M 245–261 (253), F 264, U 256, 258; bill, M 40–44 (42), F 36, U 37, 34.1; tarsus, M 69–70 (69.5), F 78, U 68, 71 (MSNG, RMNH, ZRC); the birds in juvenile plumage are smaller. Wing formula, P5 ⫽ 4 ⫽ 3 ⫽ 2 ⱖ1 ⬎ 6 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 55 cm. Large ground-cuckoo with green back, black chin, gray throat and long tail. Face has less black than Bornean Ground-cuckoo C. radiatus, sides of neck and breast are greenish gray (gray in C. radiatus), back and wing coverts are darker green than in C. radiatus, bars on abdomen are narrower, sides of breast and vent rufous, and the bird is smaller than C. radiatus.

Voice Unknown.

Range and status Southwestern Sumatra. Occurs in foothills of the Barisan Range in the southern half of the island; its known localities are Gunung Singgalang, Padang highlands, Muara Sako, Rimbopengadang and Gunung Dempo (BirdLife International 2001). Originally described and recognized as a species, then as a subspecies of Bornean Ground-cuckoo, and recently again a distinct species (Shelley 1891, Finsch 1898, 1901; Robinson and Kloss 1923b, Peters 1940, van Marle and Voous 1988; Collar and Long 1996). Resident. Ground-cuckoos are hard to see in the forests, and although biologists looked for the birds for many years, they were not seen again until November 1997, when one was found alive in a mammal trap, photographed and released, in the Bukit Barisan Selatan NP, Tanggamus District, Lampung Province (Zetra et al. 2002). Later that day a ground-cuckoo seen running on the ground was probably the same bird. Another was reported in Bukit Rimbang-Baling Wildlife Sanctuary in Riau Province in October 2000. The observations let us take an optimistic view that they are survivors. At least a third of the montane forest and most lowland forest in Sumatra has been lost in the past century, and forest loss together with large-scale movement of human populations into its range is the most serious threat to the existence of this cuckoo. The Sumatran Ground-cuckoo is critically endangered (BirdLife International 2001).

264 Bornean Ground-cuckoo Carpococcyx radiatus

Habitat and general habits Forest of hilly areas, between 300 and 1400 m. Sites where specimens had been collected were tall forest; at Gunung Dempo tall trees with fairly open undergrowth, at Muara Sako extensive, and Rimbo Pengadang included both primary and secondary forest (Robinson and Kloss 1923b). Sumatran forest habitat is currently being degraded at an alarming rate. The 1997 sighting was in primary tropical evergreen rainforest on sloping hills c. 500 m above sea level, the understory with palms, pandanus, large ferns and rattan, and the most numerous trees Stemonurus secundiflora, Pterocymbium tubulatus, Dillenia indica, Dipterocarpus kunstleri, Xanthophyllum excelsum, Cinnamomum parthenoxylon, Adina polycephala and Pometia pin-

nata.The site is near a coffee plantation.The nominally protected site is losing ground to widespread poaching, human encroachment, and timber exploitation; both commercial and smallscale logging is ongoing. On the rich volcanic soils of these bird habitats, cinnamon plantations and cabbages are replacing forests (BirdLife International 2001, Zetra et al. 2002).

Food Insects (Finsch 1898, Robinson and Kloss 1923b).

Breeding A juvenile was taken in September (MSNG), suggesting a successful breeding in the first half of the year. No other breeding information is known.

Bornean Ground-cuckoo Carpococcyx radiatus (Temminck, 1832) Calobates radiatus Temminck, 1832, Nouveau recueil de planches coloriées d’oiseaux, livre 91, pl. 538. (Pontianak district, western Borneo). Other common names: Sunda Ground-cuckoo (part), Green-billed Ground-cuckoo ( part), Radiated Fruit-cuckoo. Monotypic.

barred buff, underparts unbarred pale rufous, flanks rufous with black bars. NESTLING: Undescribed. SOURCES: AMNH, BMNH, MCZ, MZB, RMNH, ROM, SMTD, UMMZ, USNM, ZMA, ZMB.

Description ADULT: Sexes alike, head black glossed purple, back and wing coverts dull green with purple gloss and coppery red reflections, rump dark rufous with indistinct black bars, wing and tail unbarred coppery violet, sides of neck gray, lower back and rump rufous with blackish bars; underparts, chin and throat black, upper breast gray, breast and belly to under tail coverts white finely barred black, under wing coverts dark rufous; large area of bare skin on the face around eye bluish green, iris brown or gray, or gray with a narrow brown outer circle, bill green, mouth lining pale pink, tarsi and toes green. JUVENILE: Similar to adult but crown dark brown, fluffy at base of feathers, back brown, alula black, feathers have a broad buffy brown edge, wing coverts

The species was first collected in Borneo in 1826 by French naturalist P. Diard. It was described by Temminck and figured in a color plate as Calobates radiceus. Temminck himself corrected the name to Calobates radiatus in his 1838 edition of Nouveau recueil de planches coloriées d’oiseaux in the Tableau Méthodique index,Vol. 1, p. 53.

Measurements and weights Wing, M (n ⫽ 15) 240–272 (258.9 ⫾ 10.4), F (n ⫽ 8) 242–272 (252.5 ⫾ 8.9); tail, M 290–326 (303.7 ⫾ 9.8), F 286–326 (299.3 ⫾ 12.0); bill, M 44–56 (47.5 ⫾ 3.0), F 43–53 (47.6 ⫾ 3.5); tarsus, M 78–94 (85.5 ⫾ 4.3), F 82–95 (87.9 ⫾ 5.2) (AMNH, BMNH, USNM). Weight, M (n ⫽ 1) 455, F (n ⫽ 1) 540 (MZB).

Bornean Ground-cuckoo Carpococcyx radiatus 265 Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⫽ 8 ⫽ 9 ⬎ 10.

Field characters Overall length 60 cm. A large ground cuckoo with green back, black chin and throat and gray breast, barred whitish underparts, and long tail. Juvenile is brown above and pale rufous below.

Voice Four calls are known for this ground-cuckoo: (1) the main long-distance call, a repeated lowpitched double note, “thook-torr” or “koohoo”, nearly on one pitch with the first note rising and the second note falling, first note about 0.5 kHz, second note longer rising from 0.4 to 0.5 kHz, then drops to 0.3 kHz, sounding like a lowpitched Koel; (2) a rolling one-note “torrmmm” given in response to playback of its long-distance call; (3) a sharp alarm,“ark”, given along with vertical movements of the tail, perhaps the same as the coughing call “khaaa” or “heh heh heh”, given when a bird is disturbed, an apparent alarm call; and (4) a complex call given when two birds are together, a harsh call followed by dove-like descending cooing and lamb-like bleating, perhaps a combination of calls (3), (1) and the bleat (Davison 1979, MacKinnon and Phillipps 1993, Laman et al. 1997, Holmes 1997, Long and Collar 2002, Ben King).

Range and status Borneo. Resident. Ground-cuckoos are widespread but they keep a low profile. They are wary, hard to observe, and occur at low densities (Finsch 1898, Büttikofer 1900, Davison 1979, Smythies 1981, Sheldon et al. 2001). They are known from 49 localities in Borneo, including 10 in Sabah, 15 in Sarawak, four in Brunei, and 20 in Kalimantan. In Sabah they live from sea level to 50 m in low-elevation primary forest, (Sheldon et al. 2001). Recent observations of ground-cuckoos are known from Danum Valley in Sabah, Gunung Mulu NP and Similajau NP in Sarawak and Gunung Palung NP in West Kalimantan. Most recent observations are of cuckoos responding to playbacks of their recorded calls and cuckoos

captured in snares set for galliform birds (Long and Collar 2002). During seven years of forest fieldwork and surveys by three observers in Gunung Palung NP, the ground-cuckoos were seen only six times (Laman et al. 1997). They are generally absent from fragmented patches of forest (Fogden 1976, Long and Collar 2002).Vulnerable to loss of large stands of forest and to hunting, they are considered near-threatened (BirdLife International 2001).

Habitat and general habits Forest, floodplain alluvial forest in Sarawak, lowland and hill forest in Brunei, primary lowland forest and perhaps on limestone soils in Sabah, pristine lowland riverine dipterocarp forest and low swampy forest in West Kalimantan. In Sarawak they live in dry, level forest in Gunung Mulu NP (Davison 1979) and Similajau NP (Duckworth et al. 1996). In Sabah they have been seen near Baturong Caves (Sheldon et al. 2001). Although nearly all records are in lowland forest, they may occur in upslope forests as well (Long and Collar 2002). Terrestrial, they live on the ground.They run and jump, perch on logs and in trees to call, and roost in trees. Ground-cuckoos sometimes follow feeding groups of bearded pig Sus barbatus that dig in soil; the bird snatches food items from the upturned earth (Smythies 1981, Davison 1979, Laman et al. 1997). Three of five local names of ground-cuckoos translate as “pig birds” (Banks 1935, Smythies and Davison 1999). Ground-cuckoos also follow foraging sun bear Helarctos malayanus in search of termites (Long

266 Coral-billed Ground-cuckoo Carpococcyx renauldi and Collar 2002). In threat display the birds gape widely and spread both wings and the tail (Davison 1979), much as in C. renauldi (Rinke 1999).

Food Mainly insects, including beetles, giant ants and termites; they also take fruit and are known as fruitcuckoos (Finsch 1898, Davison 1979). Captive birds take mixed vegetables, raw meat, insects, dead mice, cockroaches and fish (Beddard 1901, Long and Collar 2002).

Breeding Season unknown.The nest is undescribed, although birds have been reported to build a nest (Shelford 1916). Eggs are white, 47 ⫻ 35 mm (Schönwetter 1964), laid by a captive bird, perhaps the one kept in the London Zoo and later examined by Beddard (1901). The clutch size, incubation and nestling periods are unknown. A juvenile was seen accompanied by an adult ground-cuckoo in a party of bearded pigs (Laman et al. 1997), and this is the best evidence that these ground-cuckoos provide parental care for their young.

Coral-billed Ground-cuckoo Carpococcyx renauldi Oustalet, 1896 Carpococcyx Renauldi Oustalet, 1896, Bulletin du Museum National d’Histoire Naturelle, Paris, 2, p. 314. (Province of Quang-tri, Annam) Monotypic. Other common names: Renauld’s Groundcuckoo, Annam Ground-cuckoo.

Description ADULT: Sexes alike, head, neck and throat to upper breast black; back gray, indistinct whitish band behind the black neck, rump unbarred or vermiculated dark rufous, primaries and tail black glossed purplish, secondaries grayish green, secondaries 1–6 dull purplish near tip; below, center of lower breast to belly white, flanks and legs whitish finely vermiculated with dark gray, under tail coverts gray, under wing coverts dark rufous; bare skin around eye red and violet, iris dull orange to yellowish or pink, bill dusky red, rictal area violaceous blue, feet red. JUVENILE: Forehead and crown brown, back dark gray barred rufous, rump brown, tail unbarred black, primaries black glossed purplish, secondaries and upper wing coverts dark gray with tips rufous, face brown (gray around eye), chin to upper breast unbarred gray to rufous brown, belly gray, under tail coverts rufous finely barred with black; bare facial skin gray, iris light brown, bill blackish brown, feet blackish brown.

NESTLING: Naked at hatching, skin brown, at three or four days feather quills show on wing and body, feather sheaths dark gray; palate red with white markings, comprising two raised, broad white marks on either side of the palate and a raised U-shaped shield behind that, and an arc of small white papillae anterior to the broad white marks, tongue red with a black mark in the middle; palate colors fade by the age when the bird is independent of parental care. SOURCES: AMNH, BMNH, MCZ, RMNH, UMMZ, USNM.

History The species was discovered in 1895 by Père R. P. Renauld, a missionary in Quangtri Province, Annam (Vietnam). He sent three skins to the Paris museum, where Oustalet described the species and later gave a descriptive account (Oustalet 1899). Delacour heard that the cuckoos were being captured and reared for food, and they did well in captivity. His expedition from 1926–1927 (Delacour 1927) was the first to observe the birds in the field and watch their feeding behavior and displays, and some were taken to his aviaries for further behavior observations.

Measurements and weights Wing, M (n ⫽ 12) 260–290 (278.4 ⫾ 9.5), F (n ⫽ 7) 266–294 (278.2 ⫾ 9.7); tail, M 294–348 (325.1 ⫾ 16.2), F 310–362 (330 ⫾ 17.0); bill, M 44–49

Coral-billed Ground-cuckoo Carpococcyx renauldi 267 (46.3 ⫾ 1.7), F 43–47 (44.8 ⫾ 1.7); tarsus, M 87–95 (90.4 ⫾ 2.6), F 82–95 (89.0 ⫾ 5.2) (AMNH, BMNH, MCZ, USNM). Weight, (n ⫽ 1) 400 (captive, nearly grown fledged young 44 days after hatching) (Robiller et al. 1992). Wing formula, P5 ⫽ 4 ⫽ 3 ⫽ 2 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 65 cm. Large ground cuckoo with black head and throat and gray back, and long tail glossed violet, below white finely barred with black.

Voice Territorial call is a loud, mellow, moaning whistle, “woaaaah”, “wooaa” or “wohaaau” which rises from 0.9 kHz to 1.0 kHz then drops to 0.9 kHz and lasts 1 sec, repeated again in 2–3 sec (White 1984, Scharringa 1999, Robson 2000a, NSA). In calling, the male perches on a shrub or tree, opens and closes his wings repeatedly and nods his head up and down (Wylie and Shelton 1982). The moaning call gives the bird its Lao name of “ghost chicken” (P. Round). Another call is given in duet with one bird giving a 3-note call “whup, whooup” the other bird a rolling gargle (Lekagul and Round 1991), or with male and female alternating similar calls (Hughes 1997c). The male gives a bill clack as the last sound of “coo-cuh-clack” (Atkinson 1982), and the female gives a quiet “blirrrrrrrrr” while vibrating the body and pumping the tail (Hughes 1997c).

Range and status Southeast Asia: local in continental Thailand, Laos, Cambodia, Vietnam (north and central Annam). Resident. Scarce to locally common (Evans et al. 2000, Robson 2000a), they were considered nearthreatened by Collar et al. (1994).

Habitat and general habits Broadleaf evergreen forests (lower and middle strata), second growth, scrub, vine, dense cover on ground, rocky areas.They are terrestrial and diurnal. They are shy, run from disturbance and sometimes

fly from cover. Best seen when they come to feed in certain favored sites, year after year (Ben King, Phil Round). In Khao Yai, Thailand, ground-cuckoos come to refuse tip behind a local restaurant, feed on noodles, rice and perhaps insects. Another site is an insect-rich drainage pit below a camp latrine. In observation at feeding site, a bird stands motionless for 5-10 minutes, then leaps and jumps with spread wings into the drainage bed. In feeding, it reaches neck slowly forward and works it into litter, turning leaves with the bill. On guard near the feeding site, it takes short rapid runs at small birds on the ground, supplanting them. Ground-cuckoos are quick and agile in movement, or they walk and stalk, reaching neck forward and low, extending a leg to the front of the head, then move the body forward while it holds the head in place over the ground, in slow, stealthy slides, like a rail, toes splayed apart on the ground (RBP). At other times, several ground-cuckoos share a feeding site (Ben King). They live in lowlands to 1000 m, exceptionally as high as 1500 m (Robson 2000a). Uncommon in Thailand and Vietnam, where they are trapped for food and the bird trade (Round 1988, Eames 1996).

Food Insects, small reptiles, mammals and birds (Delacour and Jabouille 1931). Captive birds require high protein diets. The birds survive and breed successfully on a diet of pigeon grain, mice, crickets, mealworms and earthworms (Delacour 1927, Ezra 1927,Wylie and Shelton 1982, Robiller et al. 1992, Rinke 1999).

268 Crested Coua Coua cristata

Breeding May to August (Robson 2000a). In the field, the nest has not been described. In captivity, it is an open platform or cup, built of twigs and branches and lined with leaves. It measures 36 cm wide and 8 cm deep, and is built on the ground or in a tree, 1–4 m above the ground. Both sexes build the nest, incubate and feed the young. Eggs are white, 44 ⫻ 34 mm (Walters 1996), the clutch is 2–6, the eggs are laid 2–3 days apart. The eggs hatch in 18–19 days. By day 10 the young have open eyes and are partly feathered, they fledge in 18–24 days and feed themselves from 28 days onwards. Growth continues

until day 40, and the young are independent at 50–60 days, when they are driven off by the male as the parents nest again. Plumage of the young bird at 45 days is mixed juvenile and adult, plumage at 90 days is like the adult. Birds are sexually mature in a year and have lived for 10 and 16 years in captivity (Ezra 1927,Atkinson 1982,Wylie and Shelton 1982, Robiller et al. 1992, Hughes 1997c). In defending the nest, the parent spreads its wings out and over the back, the large alulae spread and extending the apparent size of the bird, presenting a large front as it faces danger (photograph in Rinke 1999).

Genus Coua Schinz, 1821 Coua Schinz, 1821, Das Thierreich eingetheilt nach dem Bau der Thiere . . . , 1, p. 661. Large cuckoos of Madagascar with long tails, long legs and often a crest. Type, by monotypy, Cuculus madagascariensis Gmelin ⫽ Cuculus gigas Boddaert (Peters 1940). The German term for couas,“Seidenkuckucke” (⫽ silk cuckoos, Appert 1980), describes the soft silky sheen of the plumage, the pastel purples, pinks and

buffs. Couas have long eye-lashes. The nostril is an open slit, lower near the base of the bill than near the tip. “Coua” is a native Malagasy name for these birds.Appert (1970) recognized two species groups, the tree couas (C. cristata, C. verreauxi and C. caerulea) and the running couas (the other species), while Milne-Edwards and Grandidier (1879) noted that C. reynaudii is also a climber.Ten species.

Crested Coua Coua cristata (Linnaeus, 1766) Cuculus cristatus Linnaeus, 1766, Systema Naturae, ed. 12, 1, p. 171. (Madagascar) Polytypic. Four subspecies. Coua cristata cristata (Linné, 1766); Coua cristata pyropyga (Grandidier, 1867); Coua cristata dumonti Delacour, 1931; Coua cristata maxima Milon, 1950.

Description ADULT: Sexes alike, head gray, crest gray, back green gray, wing slightly darker, upper tail coverts washed purple, tail long, purplish blue with feathers tipped white on all but the central tail feathers; underparts, throat gray, breast pinkish gray grading into pinkish orange on lower breast, belly white, under tail coverts ochre-buff; bare skin around eye blue, anterior to eye cobalt blue, bright purplish blue or purplish brown, behind the eye mixed

sky blue or greenish blue and brilliant green, bare skin narrowly outlined in black with feathers, eyering blue, iris brown, bill black, legs and feet black. JUVENILE: Plumage loose-webbed, head gray, back and wing coverts green grey edged rufous, breast unbarred; skin around eye mostly feathered not bare, bill pale gray with a pink base and lower mandible.The palate rosette is still apparent at independence and fades by the time of postjuvenile molt.Young at four weeks are similar in appearance to their parents, but paler and the bare skin around the eye is dull blue, not bright as in the adult (Marcordes and Rinke 2000). NESTLING: Hatchlings are naked, without a trace of down, skin purplish black, darker above; bill reddish.

Crested Coua Coua cristata 269 Palate bright pinkish red, on each side the palate fold supports a raised rosette of white, forming a raised white ring around a raised white spot, contrasting in color with the bright palate.Tongue black on dorsal surface, supporting a raised outer shield outline with the posterior corners white and the shield grading to light blue around the middle of the tongue, and a central ring linked to the outer shield by a lateral spoke on each side.The nestlings are well feathered on the back by 6–8 days (Bluntschli 1938, Appert 1980, Marcordes and Rinke 2000). SOURCES: AMNH, BMNH, FMNH, MNHN, MSNG, MVZ, ROM, SMF, UMMZ.

Subspecies Coua cristata cristata (Linné, 1766); as above, crest short; Madagascar E, N and W to Mahajanga; Coua cristata dumonti (Delacour, 1931); larger, crest longer, plumage paler, under tail coverts pale rufous, white tail tips broader than in C. c. cristata; Madagascar W from Mahajanga to Morondava; Coua cristata pyropyga (Grandidier, 1867); larger, paler, under tail coverts bright rufous, white tail tip broad as in dumonti, crest intermediate in length; SW Madagascar between Morondava and Toliara, also south to Amboasary; Coua cristata maxima (Milon, 1950); larger, darker, under tail coverts lack rufous, back and inner primaries washed blue not green, known from only one specimen, perhaps a hybrid (Goodman et al. 1997); SE Madagascar (Tolagnaro).

Measurements and weights Coua c. cristata: Wing, M (n ⫽ 6) 134–148 (139.2 ⫾ 6.1), F (n ⫽ 6) 136–152 (145.9 ⫾ 7.4); tail, M 194–208 (198.4 ⫾ 5.6), F 197–202 (198.8 ⫾ 2.3); bill, M 18.6–20.9 (19.8 ⫾ 0.9), F 17.1–20.9 (19.3 ⫾ 1.6); tarsus, M 37–42 (39.3 ⫾ 2.0), F 40–44 (41.3 ⫾ 1.5) (AMNH); C. c. dumonti: U (n ⫽ 10), Wing 132–148 (140.4), tail 181–209 (195.4); C. c. pyropyga: U (n ⫽ 10), wing 157–168 (161.2), tail 197–222 (206.8); C. c. maxima: U (n ⫽ 1), wing 175, tail 232.5 (Benson et al. 1976). Weight, C. c. pyrrhopyga, M (n ⫽ 4) 135–150 (143.7), F (n ⫽ 3) 131–150 (144.7) (FMNH); all

(n ⫽ 14) 131–168 (146.9) (Goodman and Benstead 2003). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 40–44 cm. Gray arboreal coua, white below, with a pale gray crest (small dark crest in Verreaux’s Coua) and a black outline around the blue facial skin (black outline lacking in Verreaux’s Coua C. verreauxi).

Voice Loud, clear “coy coy coy . . .” with notes distinct and decreasing in volume, a chicken-like “wukwuk-wuk” given in alarm, and coos and grunts. Birds respond to calls of their neighbors, when at dawn and dusk as many as 15 birds countercall (Rand 1936, Appert 1970, Langrand 1990, Morris and Hawkins 1998, Randrianary et al. 1997).

Range and status Madagascar. Resident.Widespread, with the widest geographic and habitat range of any coua, throughout Madagascar except in the central highlands. Common around Ampijoroa and in the south near Berenty, and remarkably common in secondary forest in certain sites (Lavauden 1937, Milon 1950, 1952, Milon et al. 1973, Morris and Hawkins 1998). In the remaining blocks of forest around Tolagnaro, where the sole specimen of C. c. maxima was taken, the bird has not been seen again and is thought to be extinct (Goodman and Wilmé 2003).

270 Verreaux’s Coua Coua verreauxi

Habitat and general habits Primary and secondary forest, savanna, brushland, littoral forest and calcareous hills of southwest coast, palms, and mangroves, mainly in deciduous forest, uncommon in evergreen forest in the east; sea level to 1000 m. Arboreal, they feed in canopy foliage of the forest, move from tree to tree with long glides; they also feed on the ground, and birds and in mixed-species foraging groups. Often seen in the morning hours sunning on exposed branches with their wings drooped to absorb solar radiation (Rand 1936, Langrand 1990, Urano et al. 1994, Morris and Hawkins 1998, Goodman and Benstead 2003).

Food Large insects (caterpillars, grasshoppers and crickets (Acrididae, Euschmidtidae, Gryllidae, Pyrgomorphidae), beetles, cicadas, Phasmatodea (Phyllididae)); snails, chameleons, geckos, berries, fruits, seeds, tree flower buds; takes gum from trees (Rand 1936, Benson et al. 1976, Charles-Dominique 1976, Goodman et al. 1997). In captivity, it breeds when maintained on a diet of dog pellets, crickets and grasshoppers, with cooked cow meat and heart, and takes fruit especially mangos (Marcodes and Rinke 2000).

Displays and breeding behavior Monogamous, couas occur in pairs and both parents feed the young (Appert 1970). In captiv-

ity, a female sometimes mates with two males in rapid succession, even when another female is present on her own nest (Marcodes and Rinke 2000).

Breeding Nestbuilding is seen in November and December, also in April and May; oviduct eggs are known in October and March, fledglings are seen from late October to early February, and in June a fledged young was fed by both parents (Rand 1936, Appert 1970, 1980, Benson et al. 1976, Goodman et al. 1997). Nest is a bulky shallow bowl of twigs and rootlets, built in a tree, 4–15 m above ground, mean 8 m. At some sites both sexes build the nest, one bringing twigs and the other remaining in the nest where it arranges the material into place; at other sites only one bird is said to build the nest. Eggs are dull white, 35 ⫻ 26.5 mm, clutch 2 (1). Fledglings flutter the wings and call with a hum or hiss, and the young are attended to and fed by two adults (Appert 1970, Masuda and Ramanampamonjy 1996). Incubation and nestling periods are unknown. Like other coua species, when the nestlings beg they display the palate with open mouth, vibrate the large alula and give hissing calls. The birds have bred successfully in captivity when reared on a diet of dog pellets, crickets, and grasshoppers (Marcodes and Rinke 2000).

Verreaux’s Coua Coua verreauxi Grandidier, 1867 Coua verreauxi, Grandidier, 1867, Revue et Magasin de Zoologie Pure et Appliquée, (2), 19, p. 86. (Cape Sainte-Marie, Madagascar) Other common names: Southern Crested Coua.

Description ADULT: Sexes alike, upperparts gray, crest gray with dark tip, crest held backward, upright or forward, wing slightly darker, gray with a greenish tinge, tail dark gray with tinges of blue and purple and with broad white terminal band on all but the central tail feathers, head green-gray; underparts whitish, gray on

throat and breast, and breast white, sometimes buff on sides, the tail from below has the outer rectrices T3–5 with broad white band (42 mm or longer),T2 with a short white band (24 mm); bare skin around eye bicolored, ultramarine blue around and in front of eye and bright sky blue behind the eye, no black outline around the bare skin, eye-ring blue, iris brown to red, bill small and black, legs and feet black. JUVENILE: Similar to adult but crest short, and tail feathers narrower and more pointed (not broadly rounded as in adult) and with shorter white band

Blue Coua Coua caerulea 271 (e.g., T4 is 32–36 mm in juvenile, vs 42–52 mm in adult); bill pale. NESTLING: Undescribed. SOURCES: AMNH, BMNH, MCZ.

Measurements Wing, M (n ⫽ 9) 129–137 (133.2 ⫾ 3.1), F (n ⫽ 6) 128–132 (130.3 ⫾ 1.6); tail, M 173–195 (187.9 ⫾ 3.7), F 182–198 (189.4 ⫾ 4.8); bill, M 15.2–18.4 (16.5 ⫾ 1.0), F 16.4–17.3 (16.9 ⫾ 0.4); tarsus, M 38.3–41.1 (38.3 ⫾ 1.6), F 31.7–42.7 (37.8 ⫾ 4.6) (AMNH, BMNH, MCZ). Wing formula, P8 ⫽ 7–6 ⬎ 9 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 34 cm. Small arboreal coua with a small dark spiked crest, a pale blue spot behind the eye, and no outline of black around the bare skin of the face.

Voice Loud descending series of squawking notes, often given at dusk, “crick-crick-crick-corick-corick”, higher pitched and more rasping than the corresponding “coy coy” call of Crested Coua, also a loud growl “quark quark” followed by soft “coo coo,” descending in pitch (Rand 1936, Langrand 1990, Morris and Hawkins 1998).

Toliara. Resident. The species range is complementary to that of Coua cristata; with C. verreauxi in drier habitat. Near-threatened (Collar et al. 1994).

Habitat and general habits Subarid thorn scrub (Euphorbia and Didierea brush) and subdesert on sandy or calcareous soil; sea level to 200 m (Rand 1936, Benson et al. 1976, Langrand 1990, Morris and Hawkins 1998). These couas are mainly arboreal.

Food Insects, gekkos and small chameleons, also fruit of Cassia, taken mainly in trees (Rand 1936, Benson et al. 1976, Goodman and Benstead 2003).

Range and status SW Madagascar between Onilahy and Menarandra rivers, east of Menarandra, in coral rag scrub near

Breeding Unknown.

Blue Coua Coua caerulea (Linnaeus, 1766) Cuculus caeruleus Linnaeus, 1766, Systema Naturae, ed. 12, 1, p. 171. (Madagascar) Monotypic.

around eye blue (sometimes bicolored with anterior to eye purplish blue, behind eye cobalt blue), the bare skin around eye narrowly outlined by black feathers, iris brown, bill black, feet black.

Description ADULT: Sexes alike, upperparts and underparts dark blue, not glossy, wing and tail dark blue with violet sheen, tail with no white tip; bare skin

JUVENILE: Back and lower belly sooty black, wing dull blue, tail lacks violet; skin around eye feathered not bare, bill and feet black.

272 Blue Coua Coua caerulea NESTLING: Undescribed. SOURCES: AMNH, BMNH, CUM, FMNH, ROM, SMF, UMMZ, USNM.

Measurements and weights Wing, M (n ⫽ 6) 182–206 (195.2 ⫾ 8.0), F (n ⫽ 6) 194–213 (199.5 ⫾ 7.3); tail, M 238–276 (253.7 ⫾ 13.9), F 248–273 (258.5 ⫾ 9.5); bill, M 21.6–23.1 (22.4 ⫾ 0.6), F 19.4–25.0 (22.6 ⫾ 1.9); tarsus, M 52–57 (55.0 ⫾ 1.8), F 51–59 (54.8 ⫾ 3.1) (AMNH). Weight, M (n ⫽ 3) 225–257 (235.5), F (n ⫽ 2) 240–268 (254) (FMNH, UMMZ). Wing formula, P6 ⫽ 5 ⬎ 4 ⬎ 7 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 48 cm. An arboreal all-blue coua, with blue-violet wings and tail.

Voice Brief trilled “brrreee-ee” increasing in volume, also loud series of “coy coy coy coy” decreasing in intensity, with low pitch; also grunt “kroo kroo” (Rand 1936, Langrand 1990, Randrianary et al. 1997).

Range and status Madagascar, in most of the east and northwest. Resident, with local seasonal movements between littoral forest and nearby humid forest on laterite soils (Goodman et al. 1997). Common in suitable habitat through their range, mainly in eastern evergreen forest, seen in PN de Ranomafana, also locally in the north and northwest including Île Sainte-Marie. They are hunted and their forest habitat has become degraded (Benson et al. 1976, Goodman 1993, Morris and Hawkins 1998). Ranges of the arboreal C. caerulea and Crested Coua C. cristata are largely exclusive (Goodman et al. 1997).

Habitat and general habits Primary forest, second growth and dense clove plantations, local in deciduous forest, littoral forest, mangroves, from treetops to undergrowth, mainly in mid-strata.They occur from sea level to 1800 m; most are between 440 to 1200 m. In the PN d’Andohahela, they occur from 440 to 1875 m and are most numerous around 810 m (Hawkins and Goodman 1999); at PN de Marojejy from 450 to

1625 m, rarely higher (Goodman et al. 2000). An arboreal coua, seen when moving from tree to tree or crossing an opening, flying from higher perch and gliding lower to the next site. Slow and deliberate in movement in the trees and clumsy in flight (Rand 1936). Active from forest canopy to the undergrowth, and they run on the ground. Blue Couas, usually single birds, follow foraging troops of lemurs Eulemur fulvus (Goodman et al. 1997).

Food Insects (cicadas, locusts, crickets, katydids, bees, beetles, caterpillars, large homoptera), small reptiles (chameleons), frogs, crabs, fruits, flower buds of tree Symphonia, takes gum resins from Sloanea rhodantha trees in humid forests, the gums being rich in polysaccharides (Rand 1936, Charles-Dominique 1976, Goodman et al. 1997).

Displays and breeding behavior In courtship feeding, a bird held a frog, then opened its wings and fanned its tail, and gave a bow to the other bird (Goodman et al. 1997).

Breeding Mainly in the rains from July to December (Rand 1936, Langrand 1990). Nest is either covered above (Langrand 1990) or an open bowl-shaped clump, built of interlaced dry plant material, in dense foliage, 3.5–10 m above ground. Eggs are white, 37 ⫻ 28.5 mm, clutch 1 (Schönwetter 1964, Hawkins et al. 1998, Morris and Hawkins 1998). Incubation and nestling periods are unknown.The adult feeds insects to the young.

Red-capped Coua Coua ruficeps 273

Red-capped Coua Coua ruficeps G. R. Gray, 1846 Coua ruficeps G. R. Gray, 1846, The Genera of Birds, 2, p. 454, col. pl. cxv. (Madagascar and the eastern side of Africa ⫽ Madagascar) Polytypic. Two subspecies. Coua ruficeps ruficeps (G. R. Gray, 1846); Coua ruficeps olivaceiceps (Sharpe, 1873).

Description ADULT: Sexes alike, crown rufous or light brown green, upperparts and wing uniform light brown with green wash, upper tail coverts rufous brown, tail dark with purple cast, tail feathers broad, T1 and T2 uniform brown, T3 to T5 black tipped white; underparts, throat buffy white, breast light purplish, belly buff, center of belly whitish, under tail coverts rufous-buff; bare skin around eye indigo blue outlined in black feathers, the outline thicker behind and below the eye, the black feathers forming a band around the nape, iris dark brown, bill black, feet black. JUVENILE: Plumage dull, crown rufous buff with black feather edges, back greenish brown barred with blackish subterminal band and light brown feather edges, wing barred with gray to buff tips on coverts, primaries and secondaries black with buff terminal band, tail feathers narrow and pointed,T3 to T5 tipped whitish, breast rufous with dark gray bars; bill pale, color changes from reddish to black. NESTLING: Hatchlings naked, without a trace of down, skin purplish black, bill reddish, eyes closed. Palate bright pinkish red, on each side the palate fold supports a raised rosette of white, forming a raised white ring around a white central raised spot, contrasting in color with the bright palate.Tongue black on dorsal surface, supporting a raised shield outline, a ring around a central boss, linked to the white edge of the shield by four radial spokes. In older nestlings the white shield regresses and the red spaces enlarge, losing the anterior spoke (Berger and Lunk 1954, Appert 1967, 1970, 1980). SOURCES: AMNH, BMNH, FMNH, MVZ, ROM, SMF, UMMZ.

Subspecies Coua ruficeps ruficeps (G. R. Gray, 1846); adult, crown bright rufous; Madagascar,W and S from Mahajanga to near Morondava; Coua ruficeps olivaceiceps (Sharpe, 1873); adult, crown light brown green, upperparts and underparts paler than in C. r. ruficeps; juvenile, crown blackish gray slightly barred buff at tips of feathers, back barred, breast barred gray on white, belly white, under tail coverts gray, wing coverts with subterminal black band and buff tips, bill pale pink at base; south from Morondava to southern Madagascar.

Measurements and weights C. r. olivaceiceps: Wing, M (n ⫽ 6) 164–172 (167.5 ⫾ 2.8), F (n ⫽ 7) 161–173 (167.9 ⫾ 4.9); tail, M 214–240 (225.7 ⫾ 9.5), F 222–242 (234.2 ⫾ 6.7); bill, M 21.0–22.5 (22.0 ⫾ 0.7), F 20.4–23.8 (22.5 ⫾ 1.2); tarsus, M 54–59 (55.8 ⫾ 1.9), F 53–58 (55.4 ⫾ 1.6) (AMNH); C. r. ruficeps: Wing, M (n ⫽ 6) 161–168 (166.0 ⫾ 2.9), F (n ⫽ 3) 154–167 (160); tail, M 233–252 (239.8 ⫾ 12.4), F 231–243 (236); bill, M 22.7–24.6 (23.4 ⫾ 1.0), F 21.7–22.8 (22.4); tarsus, M 51–55.9 (53.3 ⫾ 2.1), F 48.5–53.5 (50.8) (AMNH, BMNH). Weight, C. r. ruficeps M (n ⫽ 1) 202 (BMNH); C. r. olivaceiceps U (n ⫽ 1) 182 (Goodman and Benstead 2003). Wing formula, P5 ⫽ 4 ⫽ 3 ⬎ 6 ⬎ 2 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 42 cm. Terrestrial coua with long legs, long neck and small head, slender bill, long tail and slim body which give the bird a characteristic shape, and the thick black outline behind the bare blue face is distinctive.

Voice A loud call “coy coy coy coy” rising up the scale unlike that of other couas, a loud “hug yew yew yew kuh kuh” with the last two notes lower; also grunts (Appert 1970, Langrand 1990, Morris and Hawkins 1998).

274 Red-fronted Coua Coua reynaudii forest areas (Urano et al. 1994).They live alone or in pairs (Delacour 1932,Appert 1967, 1970, Morris and Hawkins 1998). Size of territory or home range of nesting pair, 1–4 ha (Masuda and Ramanampamonjy 1996). They sunbathe on the forest floor, drooping their wings to absorb the heat, and they visit forest camp sites and feed on discorded rice (Goodman and Benstead 2003).

Food Insects (orthoptera, beetles); fruits, seeds, and rice (Milne-Edwards and Grandidier 1879, Rand 1936).

Range and status Madagascar in the west, southwest and south to Ankaoabo,Toliara,Ampotaka and Berenty. Resident. This is the only common terrestrial coua in the seasonally dry woodlands.

Habitat and general habits Dry deciduous forest, thorn scrub, spiny desert, second growth, degraded open woodland, gallery forest on edge of rivers and in forested bottomlands, in coral rag scrub near Toliara and arid habitat near Berenty; sea level to 900 m. Terrestrial, walk on the ground, usually run to escape danger, fly short distances, and perch in trees to call and to sun in the early morning. They also walk along branches. On the ground they usually walk, but when disturbed they move quickly in a jumping run (Berger 1960). They feed mainly on the ground in open

Breeding Season: October to January, females have enlarged ovary in February and April (Rand 1936, Appert 1970, 1980, Goodman et al. 1997; SMF). The nest is a shallow bowl of thin branches, bark and creepers, 15 ⫻ 25 cm and 5–12 cm thick, built in a tree 2–6 m above ground, with 19 of 37 nest sites found in the canopy and most others in the first fork of a tree or in a bush. Eggs are white, with a slightly chalky outer layer, the inner layer tinged dull blue, 34.4 ⫻ 28 mm, clutch most often 2 (1–3). Incubation period is unknown, and the nestling period is 12 days. Day-old nestlings beg in upright posture, neck stretched upward, mouth held open exposing the patterned palate, wings held straight out at the side (Appert 1970). Most nests are lost due to predation (Milon 1952, Appert 1967, 1970, 1980, Langrand 1990, Masuda and Ramanampamonjy 1996).

Red-fronted Coua Coua reynaudii Pucheran, 1845 Coua reynaudii Pucheran, 1845, Revue et Magasin de Zoologie Pure et Appliquee, 1845, p. 51. (Madagascar) Other common names: Reynaud’s Coua. Monotypic.

Description ADULT: Sexes alike, crown bright rufous, back and wing coverts dark olive green, wing green with blue wash, tail long,T1 with green gloss and T2 with blue gloss and dark tip, face blackish; underparts dark gray, darker on lower belly and under tail coverts; bare skin

around eye bicolored, ultramarine blue around and in front of eye and sky blue behind the eye, eye-ring blue, iris brown, bill black, legs and feet dark gray. JUVENILE: Crown slightly dull rufous brown, feathers loose-webbed with gray base, nape dull brown, back dull olive brown, rump rufous with the fluffy base of the feathers black, wing coverts and flight feathers tipped rufous with black subterminal band, tail dull green-brown T1, outer T2 to T5 glossed blue; underparts, throat gray, breast rufous-buff,

Red-fronted Coua Coua reynaudii 275 lower breast and belly slate brown; facial skin dull, bill yellow. NESTLING: Undescribed. SOURCES: AMNH, FMNH, ROM, UMMZ.

Measurements and weights Wing, M (n ⫽ 7) 129–140 (136.9 ⫾ 6.0), F (n ⫽ 7) 131–142 (139.1 ⫾ 3.8); tail, M 200–223 (210.1 ⫾ 6.8), F 198–222 (207.1 ⫾ 6.6); bill, M 18.8–21.0 (19.5 ⫾ 0.8), F 17.3–21.4 (19.9 ⫾ 1.4); tarsus, M 41–47 (44.0 ⫾ 2.4), F 42–46 (44.6 ⫾ 1.5) (AMNH). Weight, M (n ⫽ 2) 128–151 (139.5), F (n ⫽ 1) 163 (FMNH, UMMZ); all (n ⫽ 5) 128–175 (153.2) (Goodman and Benstead 2003). Wing formula, P6 ⫽ 5 ⫽ 4 ⬎ 7 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 38 cm. A terrestrial coua with dark plumage, no reddish color below, a rufous crown and a pale blue skin behind the eye.

Mainly terrestrial, they move slowly, walk on the forest floor and on low herbs, in forest clearings and on trails near vine tangles, move and feed on low branches and creepers, walk up sloping trunks and perch in trees (Rand 1936, Goodman et al. 1997, 2000, Hawkins et al. 1998, Morris and Hawkins 1998, Hawkins and Goodman 1999).

Food

Brief, raucous, plaintive “koo-ah” repeated several times, a sharp “jick”, and a long guttural rattle or chatter that increases in amplitude (Rand 1936, Langrand 1990, Morris and Hawkins 1998).

Insects, including beetles (Cerambycidae, Curculionidae, Elateridae, Scarabaeidae), grasshoppers (Euschmidtiidae, Tetrigidae), locusts, cicadas, stick insects (Phasmatodea), caterpillars; centipedes and spiders (Areneae) (Rand 1936, Goodman et al. 1997, Hawkins et al. 1998), also fruits and seeds (vs Coua serriana, which are mainly frugivorous).

Range and status

Displays and breeding behavior

Madagascar (E, NW). Resident. Fairly common, broadly distributed, reported in more than 30 localities, they are most numerous at higher elevations, in forest on Masoala Peninsula near Ambanizana, at Perinet-Analamazaotra and at PN de Ranomafana (Morris and Hawkins 1998).

Courtship feeding was observed when one coua accompanied another as it searched for nesting materials, then offered it a caterpillar and attempted to mate (Goodman et al. 1997).

Voice

Habitat and general habits Humid forest, dense vegetation in undisturbed rain forest and thick second growth, in tangles of fallen trees and branches, ranging into brush areas and clearings, also in dry forest in the west. Terrestrial. They occur from sea level to 2500 m, and are the most widespread coua in dwelling altitude; at PN de Marojejy in northeastern Madagascar occur from 450 to 1875 m. Unlike most other couas, they have not been seen to feed in mixed-species flocks.

Breeding August to January (Rand 1936). Nest is a shallow bowl of dry stalks, palm fibers and large leaves, 2–7 m above ground in the base of first leaves of pandanus or fern, or in a clump of sticks, or under a fallen tree near the ground where the site is covered in a net of roots and scrambling bamboo.The nest is about 19 cm in diameter, 9 cm high with wall of 5 cm, and an inside bowl of 10 ⫻ 6 cm. Eggs are chalky, dull white, 36 ⫻ 28 mm, clutch 2 (Benson et al. 1976, Hawkins et al. 1998). Incubation and nestling periods are unknown.

276 Coquerel’s Coua Coua coquereli

Coquerel’s Coua Coua coquereli Grandidier, 1867 Coua coquereli Grandidier, 1867, Revue et Magasin de Zoologie Pure et Appliquée (2), 19, p. 86, 391. (Morondava, Madagascar) Monotypic.

Description ADULT: Sexes alike, upperparts and wing olive green with dull bronze wash, tail dark bronze, face black; underparts, throat and breast whitish to tan, lower breast purplish buff, belly and under tail coverts olive gray, underside of tail blackish tipped white (T1 central tail feathers without white tips); bare skin around eye tricolored, sky-blue above the eye, darker blue around and behind the eye and opalescent pinkish-lilac behind the blue skin, the bare skin outlined by black feathers, eye-ring blue, bill black, legs and feet black. JUVENILE: Crown gray, plumage duller than in adult, wing coverts brown with buff edge, primaries and secondaries tipped buff with subterminal black band, tail with rectrices as in adult; underparts whitish barred darker gray, skin around eye slightly feathered, dull blue, face lacking black outline feathers, bare skin duller blue than in adult, iris dark brown, bill pale flesh, culmen sepia, legs and feet gray. NESTLING: At hatching, nestlings are naked, without a trace of down, skin purplish black; bill light reddish. The palate is bright pinkish red; on each side the palate fold supports a raised rosette of white, forming a raised thin white ring around a small white central ring, contrasting in color with the bright palate.Tongue is bright red with a black dorsal surface supporting a raised shield of bright pale blue outlining the edge of the shield and a central blue ring, joined to the outer shield at the base and by a radial spoke on either side (Appert 1980). SOURCES: AMNH, BMNH, MNHN, ROM, SMF, YPM.

Measurements and weights Wing, M (n ⫽ 6) 138–151 (144.0 ⫾ 4.6), F (n ⫽ 6) 143–152 (147.8 ⫾ 5.0); tail, M 200–237 (218.5 ⫾ 14.8), F 210–226 (219.7 ⫾ 7.6); bill, M 20.4–23.1 (21.8 ⫾ 0.9), F 20.0–22.3 (21.2 ⫾ 1.0); tarsus, M 47–52 (48.2 ⫾ 1.9), F 43–50 (46.3 ⫾ 2.4) (AMNH). Weight, U juvenile (n ⫽ 1) 135 (BMNH). Wing formula, P6 ⫽ 5 ⫽ 4 ⫽ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 42 cm. Medium-sized terrestrial coua with whitish throat, rufous breast and dark underparts. Differs from the larger Giant Coucal C. gigas by its more compact appearance and shorter bill, from the Running Coua C. cursor by its whitish (not pinkish buff ) throat and darker under tail coverts, and from the Red-capped C. ruficeps by its rufous belly, dark under tail coverts and all-black bill.

Voice Loud, clear series of two to four whistles, then a low soft rolling growl “weerweer-weer grull”, or inserts a couple of higher notes in a longer “weerweer-weer-wink-wink grull”; also grunts and growls, and a high pitched “ayoo-ew”. Like other couas, when one bird calls, its neighbors call as well (Appert 1970, Langrand 1990, Morris and Hawkins 1998).

Range and status Madagascar. Resident. Fairly common in north and northwest Madagascar, notably at Ampijoroa and Kirindy Forest, also Zombitse. They are subject to trapping and hunting and are shy in many areas. They occur alongside Coua gigas and C. ruficeps, and are replaced south of their range by C. cursor (Milon 1952, Benson et al. 1976, Morris and Hawkins 1998).

Running Coua Coua cursor 277

Food Insects, grasshoppers and caterpillars, also moths, beetles; spiders, berries, fruit and seeds (Rand 1936, Benson et al. 1976).

Displays and breeding behavior Monogamous, the couas occur in pairs (Appert 1970).

Breeding

Habitat and general habits Dry forest and evergreen humid forest with limited ground cover, uncommon in second growth, ranges into edge of southwestern sub-desert; sea level to 1000 m. Feeds on trails on forest floor, in carpet of dead leaves, and in forest middle layers 1–5 m above ground. A terrestrial coua, usually unhurried in movement, can fly when disturbed but more often run on the forest floor (Rand 1936, Langrand 1990, Urano et al. 1994, Morris and Hawkins 1998).

Females are in laying condition from January to March (Rand 1936), nests are active from November to January (Appert 1970, 1980, Masuda and Ramanampamonjy 1996). Nest is a bowl of twigs, small branches, petioles and bark, 16 cm across, 13 cm high with a cup 5 cm deep. Nest sites are 3–8 m above ground (mean, 5 m), built in vines either in the canopy or on the trunk of a tree (Masuda and Ramanampamonjy 1996). Eggs are dull white, 33.5 ⫻ 25 mm, clutch is 2–3 (Appert 1970, 1980, Langrand 1990). Incubation and nestling periods are unknown. Nestlings beg in upright posture, neck stretched upward, mouth held open exposing the patterned palate, wings held straight out at the side (Appert 1980).

Running Coua Coua cursor Grandidier, 1867 Coua cursor Grandidier, 1867, Revue et Magasin de Zoologie Pure et Appliquée (2), 19, p. 86, 391. (Cape Sainte-Marie and Machikora, Madagascar) Monotypic.

Description ADULT: Sexes alike, upperparts pale gray-bronze, wing green-gray, tail T1 darker bronze not tipped white, tail feathers T2 to T5 dark gray tipped white; underparts, throat orangish buff, sides of neck peach-orange, breast pale purplish grading to whitish belly and under tail coverts, flanks pale gray; bare skin around eye bicolored, deep ultramarine blue around and in front of eye and bright pink behind the eye, the bare skin outlined by black

feathers, eye-ring blue, iris dark red to brown, bill black, legs and feet black. JUVENILE: Plumage duller, crown pale gray, upperparts gray bronze edged buff, inner primaries and outer secondaries tipped buff, face dull peachorange, no black, eye-lashes present as in adult; bare skin around the eye dark bluish gray, iris dark brown, bill pale pinkish, changing gradually to black (Appert 1980). NESTLING: Naked at hatching, no trace of down on feathers of the older nestling. Palate bright pinkish red, on each side the palate fold supports a raised rosette of white, forming a raised white ring with

278 Running Coua Coua cursor well-developed fringes on the outside, and a central white irregular spot, contrasting in color with the red palate.Tongue black on dorsal surface, supporting a raised shield with white proximal corners and a blue outer circle and inner circle linked to it by a radial spoke on each side (Appert 1980). SOURCES: AMNH, BMNH, FMNH, MVZ, USNM.

Measurements and weights Wing, M (n ⫽ 7) 125–135 (129.8 ⫾ 3.8), F (n ⫽ 6) 130–142 (134.7 ⫾ 5.4); tail, M 180–200 (190.5 ⫾ 8.3), F 182–223 (200.5 ⫾ 15.2); bill, M 19.0–21.7 (21.0 ⫾ 1.5), F 20.2–20.5 (20.4); tarsus, M 40–45 (41.7 ⫾ 2.0), F 40–43 (42.0) (AMNH, BMNH, FMNH, USNM). Weight, F (n ⫽ 1) 118 (FMNH). Wing formula, P6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 7 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 34 cm. A pale, slim, long-legged terrestrial coua. The bare bright pink skin behind the eye, the peach-orange neck and the pale belly are distinctive. Similar in appearance to Coquerel’s Coua C. coquereli, but the throat is pinkish buff not white and the belly and under tail coverts are pale not dark.

Voice Loud series of six or seven gruff whistles, the first three or four are similar high pitched notes, the later notes are shorter, quieter and more muffled,“gweerrgweerr-gweerr gulk-gulk-gulk”, also grunts. Calls are similar to those of Coua coquereli (Appert 1970, Langrand 1990, Morris and Hawkins 1998).

Range and status Madagascar, from the west coast to the south. Eastern limits of range at lower western slopes of the Anosyenne Mts, also in transitional forest near Mahamavo.“Machikora” type locality has not been traced. Resident. Uncommon. They occur in the southwest and south of Madagascar, in spiny desert around Ifaty and in spiny scrub south and east of Toliara, and by Lake Anony near Berenty. It is the southern counterpart of the closely related

Coquerel’s Coua C. coquereli. This is the least numerous of the four Coua species near Mananara River in Parcel II of the PN d’Andohahela (Goodman et al. 1997, Morris and Hawkins 1998).

Habitat and general habits Subarid thorn scrub, spiny desert and dry woodlands with bare ground and an absence of ground cover, in areas of sub-desert brush and calcareous plateau with low forest bush, but not in areas of only Euphorbia bush and not in gallery forest; sea level to 200 m. Terrestrial, feed mainly on the ground, walk on the ground, and hop, and can fly with a heavy flight but more often they escape on foot, running swiftly from danger. Secretive, they live alone or in pairs (Rand 1936, Benson et al. 1976, Langrand 1990, Goodman et al. 1997).

Food Spiders, insects, including beetles (Curculionidae), cicadas (Cicadidae), ants (Formicidae), and plant materials; feed mainly on the ground (Rand 1936, Goodman et al. 1997).

Breeding In the rains; the couas lay in October and November, and young juveniles are seen in February and March (Rand 1936, Appert 1970, 1980, Langrand 1990, Goodman et al. 1997; FMNH). Nest is a bowl of twigs and bark, lined with leafstalks, in a dense bush c. 2 m above ground or in the fork of a tree. Eggs are whitish, 34 ⫻ 23 mm, clutch 2. Incubation and nestling periods are unknown.

Giant Coua Coua gigas 279

Giant Coua Coua gigas Boddaert, 1783 Coua gigas Boddaert, 1783, Table des Planches enluminéez d’histoire naturelle de Daubenton, p. 50. (Madagascar) Monotypic.

Description ADULT: Sexes alike, crown dark brown, nape, back and wing olive gray with dull bronze gloss, tail glossy bronze-black with white tips (T1 central tail feathers without white tips); underparts, throat and upper breast white, lower breast tan, belly rufous to black, under tail coverts black; bare skin around eye is tricolored with bright greenish blue above, pink to purplish below and behind eye, grayish blue anterior to eye, the bare skin outlined by black feathers, eye-ring blue, iris brown to red-brown, bill black, legs and feet black. JUVENILE: Plumage duller, less bright, wing coverts olive gray with fawn spots, primaries and secondaries olive gray with fawn tips and a black subterminal band, tail less glossy than in adult; bare skin around eye dull blue, bill pale. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, MCZ, MVZ, ROM, SMF, USNM.

Measurements and weights Wing, M (n ⫽ 6) 214–230 (220.7 ⫾ 5.9), F (n ⫽ 6) 206–220 (213.5 ⫾ 5.3); tail, M 288–318 (302.2 ⫾ 10.5), F 292–310 (305.7 ⫾ 11.1); bill, M 27–32.6 (29.9 ⫾ 2.5), F 28–29.5 (28.1 ⫾ 0.2); tarsus, M 59–68 (64.9 ⫾ 3.8), F 62–69 (66.2 ⫾ 2.6) (AMNH, FMNH, USNM). Weight, M (n ⫽ 2) 410, 415 (412.5), F (n ⫽ 1) 420 (FMNH). Wing formula, P6 ⫽ 5 ⬎ 4 ⫽ 3 ⬎ 7 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 58–62 cm. Large terrestrial coua, distinguished by a tricolored bare face with pink

behind the eye, long tarsi and large size. See Coua coquereli for similar species.

Voice Calls on the ground or in a low tree, a deep “wok wok wok . . .”, a guttural “ayoo-ew”, a resonant “kookookookoogogo” with the last notes lower, and short grunts (Rand 1936, Appert 1970, Langrand 1990).

Range and status Madagascar S and W, north to Betsiboka River. Resident. Fairly common in spiny forest and gallery forest. Common in west, less common south of its range. They persist in small forest fragments a few hectares in area, and they walk between forest patches in disturbed areas. In Berenty, the densest populations are in tall closed-canopy gallery forest where leaf litter is thick; also at Zombitse and in Kirindy Forest north of Morondava. The birds coexist with C. coquereli, C. ruficeps and C. cursor. Uncommon in areas where they are trapped and hunted. One was caught, constricted and eaten by a large boa Acrantophis dumerilii (Appert 1970, Milon et al. 1973, Goodman et al. 1997, Morris and Hawkins 1998).

Habitat and general habits Deciduous forest, thorn scrub in areas of calcium rich soil, forest and brush on sand, gallery forest, littoral forest, large trees with sparse understory, generally

280 Snail-eating Coua Coua delalandei absent in forest on lateritic soils; sea level to 1250 m. Terrestrial, they walk on dead leaves on forest floor, scratch and scrape in leaf litter for food, run quickly in pursuit of prey, changing direction at a sharp angle, pushing off trees with feet and legs, and leap high to catch prey in flight; they ruffle their feathers. In the early morning they visit sunny spots on the forest floor where they droop their wings to absorb heat (Rand 1936,Appert 1970, Goodman et al. 1997).

Food Insects and other terrestrial arthropods, including millipedes (Diplopoda), beetles (Carabidae, Curculionidae, Scarabaeidae, Tenebrionidae), ants (Formicidae), flies (Asilidae), lepidopteran larvae, grasshoppers; small reptiles, occasionally seeds (Benson et al. 1976).

Breeding Nestbuilding has been observed from late October to December, females were ready to lay in November and December, nests were seen with eggs in November and December, and chicks were seen in January (Rand 1936, Appert 1970, Goodman et al. 1997). Nest is built by both male and female, a bowl of twigs, bark and large leaves, lined with leaf petioles, 3–10 m above ground in trees (Acacia, Tamarindus), usually in vines and lianas in areas of dense vegetation, the nest is 25–40 cm across and 25 cm high. Eggs are dull white, 43.5 ⫻ 32 mm, clutch 3 (MilneEdwards and Grandidier 1879, Appert 1970, Goodman and Ravoavy 1993, Goodman et al. 1997). Incubation and nestling periods are unknown.

Snail-eating Coua Coua delalandei (Temminck, 1827) Coccycus Delalandei Temminck, 1827, Nouveau recueil de planches coloriées d’oiseaux, livre 74, pl. 440. (Madagascar) Monotypic.

Description ADULT: Sexes alike, dark violet head, above black with violet blue, wing blue with a violet sheen, tail long, violet blue, T1 and T2 blue throughout length, T3 to T5 broadly tipped white; underparts, throat to breast white, belly and flanks rufous, under tail coverts dark rufous; bare skin on face blue bordered with black feathers, iris dark brown or yellow, bill black, feet blue-gray, toes short. JUVENILE: Undescribed. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH (2), IRSNB (Brussels), MCZ, NMW, RMNH; also (not seen) Merseyside County Museums (Liverpool), MNHN (2), Stuttgart Museum and Tananarive Museum (Benson and Schütz 1971, Morgan 1975, Fisher 1981,Voisin and Voisin 1991, Collar et al. 1994).

Measurements Wing, U (n ⫽ 4) 217–226 (221); tail, 256–300 (287); bill from head, 35–42 (38.1); bill depth (n ⫽ 1) 17; bill width at base of mandible (n ⫽ 1) 15; tarsus, 68–80 (73.0) (AMNH, BMNH, MCZ). Wing formula, P6 ⫽ 5 ⫽ 4 ⫽ 3 ⫽ 2 ⬎ 1 ⬎ 7 ⬎ 8 ⬎ 9 ⬎ 10. Although sex of museum specimens is not indicated, Sganzin (1840) noted no difference between males and females. Ackerman (1841) described the as iris dark brown; Temminck’s color plate shows the iris to be yellow.

Field characters Overall length 56 cm. Large terrestrial coua with dark blue upperparts and white underparts with a rufous belly.

Voice A descending modulated call, heard both in the field and in the aviary where Ackerman (1841) observed the birds, after they cracked open the shell of a snail.

Red-breasted Coua Coua serriana 281 focus of hunters and monuments to extinction (Milne-Edwards and Grandidier 1879, Lavauden 1932, 1937, Rand 1936, Collar and Stuart 1985, Collar et al. 1994, Goodman 1993). Extinct.

Habitat and general habits

Range and status Île Sainte-Marie (Madagascar). Ackerman (1841) watched the birds in the field and in his aviary on Île Sainte Marie (Nosy Boraha), where he was based as a military surgeon in an administrative center. Sganzin (1840) found the “casseur d’escargots” (a translation of the Malagasy name ‘famachachore’ (or ‘famac-acora’) meaning the “breaker of snails”) Cuculus Madagascariensis to be not very rare when he collected several on Sainte Marie in 1831 and 1832.The last known museum specimens were taken no later than 1850 (Greenway 1967, Schifter 1973, Collar and Stuart 1985). The coua is unknown on the Madagascar mainland.The species disappeared with deforestation. Secondary causes of loss may have included snaring and hunting, the stone anvils used by the coua perhaps becoming the

Primary rain forest, near sea level. It was a large ground-dwelling bird of the rain forest floor, feeding on mollusks, and breaking the snail shells against a stone anvil with its bill (Sganzin 1840, Ackerman 1841). The coua was agile in its movements, flew well, hopped from branch to branch in the trees and from rock to rock on the ground (Ackerman 1841).

Food “Agathines” (Ackerman 1841), large forest snails of the genus Agathina Férussac 1807, a synonym of Achatina Lamarck 1799.These snails are widely distributed, originally in tropical Africa, and presumably were introduced in Madagascar before the first recorded observations of Snail-eating Coua, which may have used other snails in earlier times. Giant forest snails are now widespread through introductions for food in other tropical regions, and have now become agricultural pests. Adult snails are often more than 120 mm in length.

Breeding Unknown.

Red-breasted Coua Coua serriana Pucheran, 1845 Coua serriana Pucheran, 1845, Revue et Magasin de Zoologie Pure et Appliquée, 1845, p. 51. (Madagascar) Monotypic.

Description ADULT: Sexes alike, upperparts, head, upper back and wing coverts dark bronze to green-brown, wing blackish brown, tail black washed blue with no white on the tips; underparts, face and throat, belly and under tail coverts black, breast deep reddish chestnut, belly dark olive gray; bare skin around eye bicolored, sky blue above and a darker

ultramarine blue in front of, behind and below the eye, the bare skin outlined by black feathers, eyering blue, iris dark red to brown, bill black, deeper in form than other couas, legs and feet dark gray. JUVENILE: Upperparts dark bronze with pale buff spots, wing coverts edged rufous, primaries and secondaries edged buff with blackish subterminal markings, tail dull black without blue wash, throat brown olive not black, breast dull chestnut, belly olive; the skin around the eye is feathered or partly defeathered dull blue, bill pale with a dark tip.

282 Red-breasted Coua Coua serriana NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, SMF, UMMZ.

Measurements and weights Wing, M (n ⫽ 6) 158–170 (164.5 ⫾ 4.8), F (n ⫽ 6) 162–173 (168.0 ⫾ 3.9); tail, M 220–242 (228.5 ⫾ 9.2), F 216–242 (232.8 ⫾ 10.1); bill, M 22.9–25.4 (23.8 ⫾ 1.0), F 21.9–24.9 (23.5 ⫾ 1.3); tarsus, M 54–59 (56.2 ⫾ 2.0), F 55–58 (56.8 ⫾ 1.2) (AMNH). Weight, M (n ⫽ 1) 298 (FMNH). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 42 cm. The most terrestrial coua. Stocky body shape, dark cap, all-dark plumage with no white marks (also in Red-fronted Coua C. reynaudii) and pale blue bare skin over the eye on a dark coua are distinctive field marks.

Voice Call, a melodic high-pitched descending whistle “tee oooo”, repeats a loud, deeper, resonant “hoor ha ha”; also gives a disyllabic “chee-guall”, the second syllable like a growl, and alarmed birds give a harsh resonant growl “eeowll” (Rand 1936, Langrand 1990, Morris and Hawkins 1998).

Habitat and general habits Evergreen humid forest, undisturbed rain forest, in the northeast and east. Occur from sea level to 1000 m, and at PN de Marojejy in northeastern Madagascar from 450 to 1250 m. They live mainly in low altitudes. Secretive, terrestrial, they are found alone or in pairs as they walk and run on the ground or on a horizontal branch.They seldom fly. They feed on fruits fallen under trees with flocks of fruit-eating starlings, parrots and bulbuls; they take insects from the ground, and they dart about to capture flying prey (Rand 1936, Benson et al. 1976, Langrand 1990, Hawkins et al. 1998, Morris and Hawkins 1998, Goodman et al. 2000). The couas sunbathe in sunny areas on the forest floor.

Food Berries and fruits ( from May to July), also insects (beetles, diptera) (Milon et al. 1973).

Range and status Madagascar: Zahamena to Sambava; Sambirano. Resident. Fairly common locally in forests on Masoala Peninsula near Ambanizana, also at PerinetAnalamazaotra. They tend to occur on more open ground and lower ground than C. reynaudii (Langrand 1990, Morris and Hawkins 1998).

Breeding Season is December (Rand 1936, Langrand 1990). Nest is a bowl of inter-twined branches, 2–4 m above ground, built in an epiphytic fern, pandanus or palm. Eggs are white, size unrecorded, clutch is 2. Incubation and nestling periods are unknown.

Raffles’s Malkoha Rhinortha chlorophaea 283

Cuculinae Genus Rhinortha Vigors, 1830 Rhinortha Vigors, 1830, Mem. Raffles, 1830, p. 671. Cuckoo of the Old World tropics with black and rufous plumage and a straight bill. Type, by monotypy, Cuculus chlorophaeus Raffles 1822. The genus

refers to the smallest malkoha and the only one with conspicuous sexual dimorphism in plumage. The name refers to the bill shape (Gr. rhis, the nose; orthos, straight). One species.

Raffles’s Malkoha Rhinortha chlorophaea (Raffles, 1822) Cuculus chlorophaeus Raffles, 1822, Transactions of the Linnean Society of London 13, p. 288 (Sumatra) Monotypic. Other common names: Little Malkoha. Other names: Phaenicophaeus chlorophaeus (Raffles).

T3, juvenile, 25 mm, adult 29–31 mm; white tip length, T5, juvenile 3 mm, adult 10–12 mm; T2, juvenile 7 mm, adult 14 mm, little difference in central rectrices T1).

Description

SOURCES: AMNH, ANSP, BMNH, CM, FMNH, MCZ, MZB, ROM, SMF, UMMZ,YPM, ZMUC, ZRC.

ADULT: Sexes distinct. Male, head, back and wing coverts rufous, wing rufous with darker rufous tip of flight feathers, rump black, tail black with fine gray bars (black and gray bars each 2 mm wide) in a washboard effect (seen at close range) and broadly tipped white; underparts, throat and breast rufous, belly olive gray, under tail coverts dark gray with rufous cast; bare skin around eye light green blue, iris dark brown, bill light green, feet green-gray. Female, head and upper back light gray, back and wing coverts rufous, rump black, wings rufous with darker rufous tip of flight feathers, tail rufous with black subterminal band and white tip; underparts light gray, belly gray washed rufous, under tail coverts dark chestnut; bare skin around eye light blue green, iris dark brown, bill light green, feet green-gray. JUVENILE: Plumage like adult’s and distinct in the sexes, male tail feathers obscurely barred, female throat rufous to buffy not gray, in both sexes the tail feathers narrow and without large white tips (width of T1, juvenile 27–28 mm, adult 30–32 mm;T2 and

NESTLING: Undescribed.

Systematic comments The original species description (Raffles 1822) is of a male, insofar as it does not mention a gray head and it describes the barred tail. Swainson (1838, p. 346) described another bird, a male by plumage: “193: Anabaenus rufescens. Entirely rufous: paler beneath: tail black, tipt with white, and crossed by numerous greyish bands: vent brown”. Although the Swainson collection at Cambridge has both male and female specimens, the female lacks an original label (Benson 1999). Blyth (1842) recognized the gray-headed birds and rufousheaded birds as congeneric from their similar shape (straight bill, short tail for a malkoha), and later (Blyth 1845) recognized them as the same species, although he thought the gray-headed birds (illustrated elsewhere as another species) were juveniles. The birds were also described as a species Anadaenus ruficauda Peale 1858, whose type series (USNM 14048, 14050) shows that he considered

284 Raffles’s Malkoha Rhinortha chlorophaea the female (which he believed to be male) a new species or perhaps males of chlorophaea (Deignan (1965). The Bornean skins collected and sexed by G. Doria, A. R. Wallace and A. Everett allowed Salvadori (1874) and Sharpe (1876) to recognize the gray-headed birds as females and the rufousheaded birds as males. Geographic variation has been reported in this malkoha, but the described forms are not distinct. Baker (1919) described R. c. fuscigularis in Borneo as smaller, the plumage with a buffy wash in females. Chasen and Kloss (1930) found that females vary in the gray or rufous wash on the throat, with all intermediates between these extremes in Borneo and in Malaya, and they questioned the distinctiveness of fuscigularis. Mayr (1938a) reported that females in Sarawak and northwest Borneo have a rufous throat, females in south and east Borneo have a gray or buffy gray throat, and females in north Borneo are a mixed population; and that nostrils were usually longer and more slit-like in the north, and shorter and rounder in the south, but no differences are apparent in a larger sample. Delacour (1947) referred to “mayri” in southern Borneo, but no bird was formally described.The variation in throat color may be due to incomplete postjuvenile molt. Females (e.g. ANSP 139388, 113544 and 139387) in Sumatra with old throat feathers buff to rufous, and new throat feathers in sheath gray have retained one or more narrow juvenile tail feathers with a little white on the tips. Ripley (1942) suggested that the birds on Batu Islands, West Sumatra, were larger than the birds in Sumatra, and named them R. c. facta, based on a sample of two birds. Similarly, R. c. bangkanus (Meyer de Schauensee 1958) was based on two birds from Bangka I, and a larger sample shows no consistent difference in size (Mees 1986).

Measurements and weights Wing, M (n ⫽ 19) 112–122 (114.6 ⫾ 2.5), F (n ⫽ 17) 113–118.5 (115.9 ⫾ 1.7); tail, M 154–183 (167.2 ⫾ 10.1), F 155–187 (176.0 ⫾ 7.0); bill, M 23–31 (27.1 ⫾ 2.3), F 24–29.5 (27.8 ⫾ 1.9); tarsus, M 21–27.7 (25.9 ⫾ 1.6), F 23–29.8 (25.7 ⫾ 2.1) (ANSP, UMMZ, ZMUC).

Weight, M (n ⫽ 14) 45–62 (53.2), F (n ⫽ 7) 41–62 (49.3) (AMNH, BMNH, MZB,Wells 1999). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⱖ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 32 cm. Small malkoha, the male with a rufous ginger head and back, barred blackish tail with black subterminal band and broad white tip, rufous throat and upper breast, and green bill; the female with a gray head and throat, and the underparts grading light gray to rufous.

Voice A querulous reedy call, 3-6 descending, cat-like notes “kiau kiau kiau …”, the first note also described as “miaow”, the later notes on the same pitch or lower]; other calls “heeah” or “heaeew” single or double, harsh strained croaking sounds, and a chirp (Hume and Davison 1878, Lekagul and Round 1991, Robson 2000a).

Range and status Southeast Asia in southern Burma (Tenasserim), and in extreme western Thailand and the Malay Peninsula, in Sumatra and Tanahmasa I in the western Sumatra islands, Bangka I, Borneo and Natuna I; formerly Singapore, and absent on most islands near Sumatra and Borneo (Smythies 1940, 1981, Ripley 1943, 1944, van Marle and Voous 1988, Lekagul and Round 1991, Holmes 1997, Webster and Fook 1997, Wells 1999). Resident. Common and widespread.

Whistling Yellowbill Ceuthmochares australis 285

Habitat and general habits Lowland secondary forest, forest edge, dipterocarp forest, dense jungle, peat swamp, heath scrub, cane brakes, old overgrown plantations (rubber, cocoa, Albizia, Eucalyptus), creepers on tree trunks, gardens; lowlands to 1000 m. Malkohas live in forest regrowth and on hill slopes, in disturbed areas after storm damage, logging, and in extensive secondary forests and old rubber plantations near forest. They creep through dense vegetation, looking like squirrels, feed in the middle story at a mean height of 15–20 m, where they search on leaves and peer under leaves for insects. Solitary or in pairs, occasionally in small groups and in mixed-species feeding flocks (Chasen 1939, King and Dickinson 1975, Wilkinson et al. 1991,Wells 1999, Sheldon et al. 2001, P. Round).

Food Insects, mainly caterpillars; butterflies, crickets and locusts, cicadas, and beetles; and phasmids

(Hume and Davison 1878, Smythies 1981, Wells 1999).

Breeding In Malay Peninsula they breed from March to May (Wells 1999), in Sumatra in January (van Marle and Voous 1988), in Borneo in October (adults with fledged young in November, Nash and Nash 1985), January (oviduct egg) (Medway, in Smythies 1981) and February (egg in nest, Sheldon et al. 2001).The nest is a shallow platform of twigs lined with green leaves, located 3–8 m up in a forest thicket, built by the male (Wells 1999, Sheldon et al. 2001). Eggs are white, 28 ⫻ 24 mm (Chasen 1939), in another set 33 ⫻ 25 mm (Baker 1934, Schönwetter 1964), perhaps not this species.The clutch is 2, occasionally 1 or 3 (Baker 1934, Chasen 1939, Sheldon et al. 1991). Incubation and nestling periods are unknown. Sometimes 3 young fledge in a brood (Nash and Nash 1985).

Genus Ceuthmochares Cabanis and Heine Ceuthmochares Cabanis and Heine, 1863, Museum Heineanum, 4 (1), p. 60. Plumage glossy green, blue or violet, wings short and rounded, tail long, broad and graduated, nostrils slit-like, bare skin around eye brightly colored, the bill short, broad and swollen. The only African malkohas. Type, by monotypy, Cuculus aereus Vieillot 1817. The name

is derived from Gr. keuthmos, a hiding place; khairo, to delight in, referring to the secretive behavior of the bird in dense forest. Formerly recognized as a single species; the birds in West and Central Africa and in East and southern Africa are morphologically distinct and they have distinct songs. Two species.

Whistling Yellowbill Ceuthmochares australis Sharpe 1873 Ceuthmochares australis Sharpe, 1873, Proceedings of the Zoological Society of London, 1873, p. 609. (Natal) Other common names: Green Malkoha (part), Blue Malkoha, Green Coucal (part). Monotypic.

Description

long and graduated, uniformly bright glossy green; crown and nape gray glossed green, throat light gray, breast gray washed olive buff, belly gray, under tail coverts blackish gray; bare skin in front of eye yellow, around and behind eye green to blue, iris red or brown, bill yellow (base of culmen black), feet black.

ADULT: Sexes alike, upperparts, head gray, back to wing coverts and rump dark gray glossed green, wing short and rounded, glossed as on back, tail

JUVENILE: Similar to adult, the upper and under wing coverts narrowly edged buff (not noticeable

286 Whistling Yellowbill Ceuthmochares australis in the field), rectrices narrower than in adult; iris brown, bill narrow, color dark horn changing with maturity to pale yellow. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, BWYO, CM, DM, FMNH, ROM, UMMZ, USNM, ZMUC, ZSMK.

Taxonomic comments and geographic variation The consistent differences in plumage and in song between Ceuthmochares from West Africa and Central Africa, and coastal East and southern Africa, indicate that these malkohas comprise two distinct species. Birds in southern part of the range average larger but not significantly so; C. a. dendrobates Clancey, 1962 is a synonym of Ceuthmochares australis Sharpe, 1873. Called “South African Coucal” (Layard 1874) and “Green Coucal” (Sclater 1930), the yellowbill malkohas Ceuthmochares are not related to the coucals Centropus. Bannerman (1933) proposed the common name “Yellowbill”.

Measurements and weights Kenya (coastal, Sokoke to Mombasa), wing, M (n ⫽ 8) 115–126 (120.5 ⫾ 4.6), F (n ⫽ 7) 117–129 (124.8 ⫾ 3.4); tail, M 190–210 (198.6 ⫾ 10.3), F189–213 (203.5 ⫾ 7.8); bill, M 25.2–30.7 (26.6 ⫾ 2.1), F 23.2–30.5 (26.4 ⫾ 3.0); tarsus, M 19.0– 28.5 (24.6 ⫾ 3.5), F 23.2–27.5 (25.2 ⫾ 1.6) (CM, FMNH); Tanzania (coastal), wing, M (n ⫽ 6) 122– 127 (123.5 ⫾ 3.2), F (n ⫽ 7) 117–129 (123.9 ⫾ 4.5); tail, M 188–210 (195.8 ⫾ 8.6), F 176–203 (186.9 ⫾ 10.3); bill, M 23–26 (24.6 ⫾ 1.0), F 23–28 (25.2 ⫾ 1.7); tarsus, M 27–29 (27.7 ⫾ 0.9), F 27–29 (27.7 ⫾ 0.9) (ZMUC); Malawi and Mozambique, wing, M (n ⫽ 6) 120–131 (125.8 ⫾ 3.9); F (n ⫽ 3) 126–138 (130.3); tail, M 210–222 (214.2 ⫾ 6.3), F 212–218 (215) (BMNH, UMMZ, ZMK); South Africa, M (n ⫽ 3) 124–131 (127); F (n ⫽ 2) 123–126 (124.5); tail, MF 190–210 (199.2) (Maclean 1985). Weights, East Africa, M (n ⫽ 17) 55–72 (61.9), F (n ⫽ 8) 52–72 (62.1) (Irwin 1988); South Africa, M (n ⫽ 4) 63.1–79.2 (69.5), F (n ⫽ 3) 60.4–72.4 (65.5) (Maclean 1985).

Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 4 ⫽ 8 ⬎ 3 ⬎ 2 ⱖ 9 ⬎ 1 ⱖ 10.

Field characters Overall length 33 cm. Arboreal cuckoo, slender, plumage gray with a glossy green back and tail, the tail very long and graduated, without white spots at the tip, and the bill yellow.

Voice In coastal East Africa, song is characterized by a series of low-pitched nasal whistles (Sokoke Forest, C. R. Barlow recordings; Mrima Hill, NSA 18022). Song begins with “si” notes that rise to 2.5 kHz in pitch, last about 0.1 sec, and repeat 6 per sec, then song continues with a series of low pitched whistles that become shorter and repeat more rapidly at the end of the song; the song lasts about 6 sec.The song is a long “si-si- . . . . tsik tsik . . . tik tik . . . teeew . . . teew . . . tew . . . wip wip . . . wipwipwip”. Most sound energy is at 1 kHz (overtones produce the nasal sound); the notes rise quickly then fall gradually in pitch from 1.2 to 0.9 kHz, and the whistles become shorter (0.35 to 0.08 sec) and repeat more rapidly (time between notes 0.2 to 0.08 sec). Except for high pitched “si” notes at the onset, the song stays on one pitch in a series of nasal whistles not a rattle of harsh “tsik” notes. In South Africa most songs are similar to songs in Kenya, but in a few the whistles lack overtones and are longer in duration, some as long as 0.6 sec (Figure 6.7). Most songs in E Tanzania and South Africa have both the long whistles and the overtones (NSA 06925, 10958, 53859-53864, 53867).

Range and status East Africa from Ethiopia and Kenya to Tanzania (including coastal islands), Malawi, eastern Zimbabwe, Mozambique and coastal eastern South Africa. Occasional in Somalia (Ash and Miskell 1998). In East Africa, widespread and locally common, with a gap in distribution between C. australis and C. aereus in inland Kenya and Tanzania where habitat is not forested (Stevenson and Fanshawe 2002) and a gap between C. australis in Malawi and C. aereus in northeastern Zambia. In southern Africa the yellowbills live in coastal habitats in Mozambique

Chattering Yellowbill Ceuthmochares aereus 287 1989, Zimmerman et al. 1996).They remain high in the canopy (8–30 m), moving by short hops as they balance with the long tail and search through tangled creepers, turning the body and seizing an insect. At times they accompany mixed species flocks of birds or squirrels that flush the insects (Irwin 1956, Zimmerman 1972, Lewis and Pomeroy 1989). Occur singly, in pairs or in small groups.

Food

and northern Natal, occasionally further south as far as Uitenhage, and inland in Zimbabwe on the Zambezi, Haroni-Rusitu and Limpopo Rivers (Niven and Niven 1966, Skead 1967, 1995, Rowan 1983, Quickelberge 1989,Vernon et al. 1990, 1997, Parker 1994, Clancey 1996, Masterson and Parkes 1998). Resident in most areas, perhaps migrant in coastal Kenya where most observations are May to September and specimens are June through December (Irwin 1988, Lewis and Pomeroy 1989, Zimmerman et al. 1996, CM, FMNH).

Habitat and general habits Forest with dense tangles, thickets, creepers, forest edge, treefall gaps, in upper canopy of old secondary evergreen forest and riverine forest, and coastal scrub.They are active in trees above ground, in forest clearings, tall dense thickets, tangled creepers and lianas. They occur in the littoral region in southern Mozambique and Natal, in the Usambara and North Pare Mts and Arusha and Kilimanjaro in Tanzania, in coastal forest in eastern Kenya, in thorn bush along Omo River in Ethiopia in arid habitat and in river thickets in southern Somalia; in the lowlands (Urban and Brown 1971, Irwin 1988, Lewis and Pomeroy

Insects, including caterpillars, beetles, grasshoppers, crickets, leaf-hoppers, winged termites, bees, spiders; tree frogs, slugs; fruit, seeds and leaves (Zimmerman 1972, Rowan 1983, Irwin 1988).

Displays and breeding behavior In courtship behavior, two birds perch and face each other, they wag their tails from side to side, then one spreads its tail. Male approaches the female with food in his bill. In courtship feeding, he mounts and feeds her as they copulate (Rowan 1983).

Breeding In southern Africa they breed from October to December (Dean 1971, Benson and Benson 1977, Rowan 1983, Maclean 1993,Vernon et al. 1997). The nest is an open platform made of a mass of sticks to which green leaves may still be attached (Stark and Sclater 1903, Rowan 1983, Maclean 1993,Tarboton 2001), built 2–5 m above ground in dense vegetation; closed nests have been reported in Natal (Tarboton 2001). Eggs are white to creamy, sometimes chalky, becoming stained in the nest, 30 ⫻ 23 mm, clutch 2(3), in Natal 1–4 (2 is most common) (Chapin 1939, James 1970, Dean 1971, Irwin 1988). Incubation and nestling periods are unknown. Both parents feed the young (Rowan 1983).

Chattering Yellowbill Ceuthmochares aereus (Vieillot, 1817) Cuculus aereus Vieillot, 1817, Nouveau Dictionnaire d’Histoire naturelle . . . , 8, p. 229. (Malimbe, Portuguese Congo ⫽ Cabinda)

Other common names: Senegal Yellowbill, Congo Yellowbill, Green Malkoha (part), Green Coucal (part).

288 Chattering Yellowbill Ceuthmochares aereus Polytypic. Two subspecies. Ceuthmochares aereus aereus (Viaillot, 1817); Ceuthmochares aereus flavirostris (Swainson, 1837).

Description ADULT: Sexes alike, crown and nape dark gray, grading on the back to dark gray glossed bronzy greenish-blue, blue or violet, tail glossed greenish blue, blue or violet; underparts gray, wing short and rounded, glossed as on back, tail long and graduated, glossy blue to greenish blue; bare skin around eye yellow, tinged green near the eye, iris red or brown, bill yellow (base of culmen black), feet black. JUVENILE: Similar to adult, rectrices narrower; iris brown, bill narrow, color dark gray changing with maturity to horn and then pale yellow. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, CM, DMNH, FMNH, MCZ, MSNM, UMMZ, USNM, ZSMK.

Subspecies and geographic variation Ceuthmochares aereus aereus (Vieillot, 1817); back and wings glossed bronzy blue or greenish-blue, tail glossed blue to greenish blue, the chin and breast light gray, belly darker gray, under tail coverts slaty gray; eastern, central and southern Nigeria and Cameroon to East-Central Africa; Ceuthmochares aereus flavirostris (Swainson, 1837); back and wings glossed violet, tail glossed bright violet, the chin, throat and breast gray, belly and under tail coverts slaty gray; Senegal and Sierra Leone to Ghana and SW Nigeria. Ceuthmochares aereus (Vieillot 1817) includes birds described as C. intermedius Sharpe 1884 and C. a. extensicaudus Clancey 1962. Sharpe (1884) described C. intermedius from Semmio (Zémio) on the Ubangi-Shari in the Central African Republic, distinguished it from C. flavirostris and C. australis, and noted the bird was intermediate in plumage color between these two forms. Although Sharpe remarked that intermedius was the same in plumage color as birds in Gabon and Cameroon, he overlooked the prior name Cuculus aereus Vieillot 1817

from Malimbe, Cabinda. Reichenow (1904) recognized intermedius as a subspecies of C. aereus (sensu strictu), Shelly (1891) listed intermedius as a synonym of C. aereus, Sclater (1930) listed intermedius as a subspecies of C. aereus, Bannerman (1933: footnote p. 135) confused intermedius with australis, and Chapin (1939) recognized C. a. intermedius as the form in the upper Congo (Zaire) and Uganda, noting it “differs but slightly” and more greenish (less bluish) than nominate C. aereus. Size varies across Central Africa with birds in the west somewhat smaller than birds in the east and southeast, in the range of “intermedius”, although size overlaps across this region. Birds of southern Nigeria vary from greenishblue to violet. Sclater (1930) and Bannerman (1933) attributed the birds of Nigeria to flavirostris not aereus; and Bannerman (1951) and later authors ascribed birds east of Lagos or east of the Niger River as aereus, and birds west of the river as flavirostris. In the specimens now available, birds west of the Niger River are violet with a dark gray throat and breast at Ilorin (n⫽2) and Ibadan (n⫽1); another at Ibadan (n⫽1) is blue with a pale gray throat and breast. Birds east of the Niger River are blue to blue-green with a pale gray throat and breast, much like birds of Cameroon. These specimens indicate a rather abrupt geographic change in Nigeria between violet and blue; no song recordings are available in this region.

Measurements and weights C. a. flavirostris, Sierra Leone, Guinea and GuineaBissau, wing, M (n ⫽ 8) 112-127 (116.3 ⫾ 4.7), F (n ⫽ 5) 117–123 (120.4 ⫾ 2.2); tail, M 184–202 (192.0 ⫾ 6.3), F (n ⫽ 5) 187–210 (200.6 ⫾ 9.8); bill, M 23–28 (25.1 ⫾ 1.6), F 23.1–28 (26.1 ⫾ 2.2); tarsus, M 24.1–31 (27.1 ⫾ 2.1), F 23.1–28 (26.1 ⫾ 2.1) (BMNH, MSNM); Liberia and Ivory Coast, wing, M (n ⫽ 6) 114–117 (115.6 ⫾ 1.1), F (n ⫽ 7) 112–122 (115.0 ⫾ 3.6); tail, M 186–207 (195.5 ⫾ 9.8); F 189–205 (198.0 ⫾ 5.7) (FMNH); C. a. aereus, Nigeria and W Cameroon, wing, M (n ⫽ 8) 111–120 (114.5 ⫾ 3.3), F (n ⫽ 9) 114–128 (119.9 ⫾ 4.4); tail, M 176–206 (190.1 ⫾ 10.9), F 172–225 (197.7 ⫾ 14.7); bill, M 23.5–26.0 (24.8 ⫾ 0.8), F 22.9–26.1 (24.0 ⫾ 1.0) (BMNH, ZSMK); Bioko, wing, M (n ⫽ 4) 103–l15 (109.3), F

Chattering Yellowbill Ceuthmochares aereus 289 (n ⫽ 4) 108–117 (113.5), tail, M 170–178 (174.7), F 172–184 (177.0) (ZSMK); Angola, wing, M (n ⫽ 6) 116–124 (118.2 ⫾ 3.2), F (n ⫽ 4) 115–123 (118.5); tail, M 178–199 (188.6 ⫾ 7.0), F 184–201 (192.8) (AMNH, MCZ); Zaire: Kasai, wing, M (n ⫽ 10) 107–121 (113.3 ⫾ 4.5), F (n ⫽ 3) 110–123 (116.7); tail, M 181–206 (193.8 ⫾ 7.8), F 188–208 (197.3) (AMNH); Kivu, wing, M (n ⫽ 11) 107.5–130 (118.8 ⫾ 7.2), F (n ⫽ 11) 115–129 (120.4 ⫾ 4.6); tail, M 143–219 (204.4 ⫾ 48.6), F 197–225 (206.8 ⫾ 10.1) (AMNH, MCZ); NE Zaire (Uele, Ituri), Central African Republic (Zémio) and Sudan (nr Yambio), wing, M (n ⫽ 9) 107.5–120 (114.1 ⫾ 4.8), F (n ⫽ 9); 110–124 (115.6 ⫾ 3.8); tail, M 175–195 (189 ⫾ 7.4), F 183–217 (205.0 ⫾ 8.2) (AMNH, FMNH, MCZ); Kivu, wing, M (n ⫽ 11) 107.5–130 (118.8 ⫾ 7.2), F (n ⫽ 11) 115–129 (120.4 ⫾ 4.6); tail, M 143–219 (204.4 ⫾ 48.6), F 197–225 (206.8 ⫾ 10.1) (AMNH, MCZ); Shaba, wing, M (n ⫽ 4) 126–127 (126.2 ⫾ 0.4), F (n ⫽ 7) 123–130 (124.6 ⫾ 2.4) (Verheyen 1953); and Uganda, wing, M (n ⫽ 16) 112–123 (117.2 ⫾ 3.4), F (n ⫽ 15) 112–126 (116.9 ⫾ 3.9); tail, M 173–210 (195.5 ⫾ 9.8), F 178–221 (206.5 ⫾ 13.8); bill, M 27–30.5 (28.6 ⫾ 1.0), F 27–31 (28.8 ⫾ 1.2); tarsus, M 26.5–31.5 (28.3 ⫾ 2.1), F 24–31.5 (27.6 ⫾ 1.8) (FMNH). Weights, Cameroon and Bioko, M (n ⫽ 3) 54–62 (59.3), F (n ⫽ 6) 62–73 (66.2) (Eisentraut 1963, 1973, ZSMK), Upemba NP, M (n ⫽ 4) 64–74 (70.2), F (n ⫽ 6) 57–80 (67.3) (Verheyen 1953), Sudan, F (n ⫽ 1) 56 (FMNH). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 4 ⫽ 8 ⬎ 3 ⬎ 2 ⱖ 9 ⬎ 1 ⱖ 10.

increasing in tempo and running into a rattle at 18 notes per sec, the song as long as 12 sec. In Cameroon, Bates (1930: 190) described it as “some queer little high-pitched notes in a voice suitable to a bird the size of a wren, which begin to slow but become so rapid as to run into a trill”, and songs I heard in 1995 at Buru in the Donga River region in SE Nigeria were similar in the high pitch of the first notes and the descent to the rattle. Song recordings in Central Africa from Cameroon to northeastern Zaire, Rwanda and western Kenya are the same as these descriptions (Figure 6.7). Song is sometimes introduced by 0.1 sec “tsik” notes at 2.5 kHz which morph into the shorter high notes “tsik” an octave higher (eastern Zaire, Walikale, Obale, west of Lake Kivu, NSA 80609; Nyungwe Forest, Rwanda, NSA 30638. cf. Dowsett-Lemaire 1990; Kakamega Forest, western Kenya, NSA 48766). In West Africa, song sometimes stays on one pitch (Bijagos Archipelago in Guinea-Bissau: C. R. Barlow recordings; N’Douci, Ivory Coast: Chappuis 2000) (Figure 6.7); however, Kelsall (1914: 216) described song in Sierra Leone as “notes in a descending chromatic scale commencing slowly and gradually increasing in rapidity”, and in Ghana (NSA 80927) in the geographic range of C.a.flavirostris, songs are similar to the songs in the Central African range of C. a. aereus. Calls, a mournful whistle “kou-lee, kou-lee” of 1 sec rising then dropping at the end; also a squirrel-like call “tsik-tsik”.

Field characters

In West Africa they occur from The Gambia and Casamance to Guinea, Sierra Leone, Liberia, Ivory Coast, Ghana and Togo, mainly on the African continent, also on the Bijagos Archipelago and on river and coastal islands of Sierra Leone (Bannerman 1933, Cheke and Walsh 1996). In southwestern Senegal, common around Oussouye in PN de Basse-Casamance (Morel and Morel 1990), in Guinea-Bissau common on the Bijagos Archipelago (C. R. Barlow), in Liberia from the coast to the northern hills (Gatter 1997), in Ivory Coast in forest and gallery woodlands (Thiollay 1985), and in Ghana in thickets of charcoaled forests near Cape Coast (RBP). In Central Africa they occur from southern Nigeria, southern Cameroon, Gabon,

Overall length 33 cm. Arboreal cuckoo, slender, plumage dark gray with a glossy violet, blue or greenish-blue back and tail, the tail very long and graduated, without white spots at the end, gray underparts, and a yellow bill. Somewhat squirrellike in movements and chittering calls.

Voice Song is a series of notes becoming shorter and the rate accelerating towards the end.The song begins with short (⬍ 0.1 sec) high notes “tsik” each with a wide frequency band loudest at 4 kHz, starting at a high pitch then dropping to a lower pitch,

Range and status

290 Chattering Yellowbill Ceuthmochares aereus Equatorial Guinea, Congo-Brazzaville, southern Central African Republic and southern Sudan, throughout Zaire, Angola, Uganda, western Kenya (Kakamega), northern and northwest Zambia, and in the Gulf of Guinea (Bioko), where common (Pérez de Val 1996). In Central African Republic they occur north of Ubangi River at Zémio, Sibut and the Dzonga Sangha Reserve, and they are common in forested areas of Gabon and Zaire (Chapin 1939, Brosset and Erard 1986, AMNH). In EastCentral Africa, they are widespread and locally common, with a gap in distribution between C. australis and C. aereus in inland Kenya and Tanzania where habitat is not forested, and between these two taxa where C. aereus are in northeastern Zambia and the more eastern C. australis are in Malawi. Resident in most areas, seasonal from November through June in isolated forests in The Gambia (Barlow et al. 1997), and perhaps migrant in coastal Nigeria (Elgood et al. 1994). Population density in NE Gabon is estimated at 15-20 pairs / km2 (Brosset and Erard 1986).

Habitat and general habits Forest with dense tangles, thickets, creepers, treefall gaps, in upper canopy of old secondary evergreen forest and riverine forest, coastal scrub and forest edge. They are active in trees, in forest clearings, tall dense thickets, tangled creepers and lianas. Mainly in the lowlands, they occur from sea level to 1600 m in western Cameroon and 2000 m in western East Africa (Urban and Brown 1971, Irwin 1988, Lewis and Pomeroy 1989, Zimmerman et al. 1996, Bowden 2001). On Mt Nimba in Liberia, they occur in Musanga-dominated second-growth thickets at 700 and 900 m (Colston and CurryLindahl 1986).They remain high in the canopy, foraging 10 to 30 m above the forest floor, moving by short hops as they balance with the long tail and search through tangled creepers, turning and seizing an insect. At times, the malkoha accompanies mixed species flocks of birds or squirrels that flush the insects, then it seizes the insect (Bates 1930, Bannerman 1933, 1951, Serle 1950b, Brosset and Erard 1986,Thomas 1991). Occur singly, in pairs or in small groups.

Food Insects, including caterpillars, beetles, grasshoppers, crickets, leaf-hoppers, winged termites, bees, spiders; tree frogs, slugs; fruit, seeds and leaves (Bates 1930, Bannerman 1933, Chapin 1939, Eisentraut 1973).

Breeding In Liberia they call mainly from August to October and March to May, young are in the care of their parents in October and January, adults molt from December to April and are in fresh plumage June to October (Coulson and Curry-Lindahl 1986, Gatter 1997). In Ivory Coast a laying female was taken in December (Bannerman 1933). In Ghana they nest in March and May (Morrison 1947, Grimes 1987), in Nigeria in June and August (Elgood et al. 1994), in western Cameroon in November (Serle 1950b), in Gabon a fledged young was with adults in January (Christy and Clarke 1994), in Uele a female was laying in July (Chapin 1939), in southern Zaire the cuckoos breed in June, December and February (Verheyen 1953), and in Angola they breed in May (Heinrich 1958, Dean 2000). The nest is an open mass of sticks to which green leaves may still attach), built in a tree, 2–5 m above ground in dense vegetation in a tree. It looks like a pile of leafy rubbish in a tree (Morrison 1947). Closed nests have been described in Uganda (van Someren 1949). Eggs are white with chalky patches, in shape rounded at both ends, 27 ⫻ 21 mm (Bannerman 1933, Chapin 1939, van Someren 1949). Incubation and nestling periods are unknown.

Sirkeer Malkoha Taccocua leschenaultii 291

Genus Taccocua Lesson, 1830 Taccocua Lesson 1830, Traité d’Ornithologie, livre 2, p. 143. Type, by monotypy, Taccocua leschenaultii Lesson 1830. The genus name is derived from the French “taco” for lizard cuckoos Saurothera (Lesson,

1830) and Coua the genus coua.The genus is characterized by non-iridescent plumage and a feathered face. One species.

Sirkeer Malkoha Taccocua leschenaultii Lesson, 1830 Taccocua Leschenaultii Lesson, 1830, Traité d’Ornithologie, livre 2, p. 144. (India [⫽ Madras]) Polytypic. Three subspecies. Taccocua leschenaultii leschenaultii Lesson, 1830; Taccocua leschenaultii sirkee J. E. Gray, 1831; Taccocua leschenaultii infuscata Blyth, 1845. Other names: Phaenicophaeus leschenaultii (Lesson, 1830). Other common names: Sirkeer Cuckoo.

Description ADULT: Sexes alike, upperparts brown with fine black shaft streaks on head and nape, wing brown, rump and long upper tail coverts brown, tail long, blackish, inner rectrices T1 dark brown without white tips, T2 to T5 with broad white tips; underparts paler, grading from whitish throat to gray breast and with shaft streaks straw-colored on throat and blackish tips on upper breast, the stiff rachi forming showy hackles, belly rufous buff, under tail coverts brown, underside of tail black with bold white tips, under wing coverts light brown; blackish skin around eye bordered above and below with a fine white feathered line, iris red to brown, bill red with yellow tip, feet dark gray. JUVENILE: More heavily streaked than adult, head, upper back and breast with broad black streaks, feathers on head and back edged buff to rufous, rectrices narrower, tail spots whitish buff, less distinct and smaller than in adult. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, FMNH, MSU, ROM, UMMZ,YPM.

Subspecies Taccocua leschenaultii sirkee ( J. E. Gray, 1831); plumage light gray, throat white, face with white superciliary line; Pakistan (Sind, Punjab) and NW India; Taccocua leschenaultii infuscata Blyth, 1845; plumage darker, throat gray; India in subHimalayan terai and duars (terai, the alluvial lowland areas between the Himalayas and the plains, formerly forested and well watered, now intensively farmed, from Uttar Pradesh east to northern West Bengal; duars, the same habitat from Jalpaiguri, Bhutan and Assam) south to 18° N; Taccocua leschenaultii leschenaultii Lesson, 1830; plumage dark brown, throat gray; southern India and Sri Lanka.

Measurements and weights India and Nepal: Wing, M (n ⫽ 15) 147–169 (157.6 ⫾ 5.6), F (n ⫽ 10) 148–163 (153.5 ⫾ 4.4); tail, M 216–262 (239.8 ⫾ 14.8), F 213–254 (235.2 ⫾ 12.7); bill, M 24.9–33.5 (29.4 ⫾ 3.1), F 27.0–30.5 (29.1 ⫾ 1.5); tarsus, M 36.5–43.5 (39.7 ⫾ 2.6), F 36.3–41.2 (39.3 ⫾ 1.8) (FMNH, MSU, UMMZ,YPM). Weight, M (n ⫽ 4) 174–249 (218.5), F (n ⫽ 2) 132–148 (140.0) (BMNH, weights recorded in ounces). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 8 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 43 cm. Long-tailed cuckoo with ashy brown plumage and fine black shaft streaks, a bushy crest and bill stout, red and yellow. Brown

292 Sirkeer Malkoha Taccocua leschenaultii plumage distinguishes it from other malkohas, female Koels Eudynamys scolopacea are spotted above and streaked and spotted below, and juvenile or first-year winter-plumage coucals Centropus sinensis and C. bengalensis are barred on the wing and tail and lack a white tail tip.

Voice Largely silent, occasional shriek “kek-kek-kekkerek-kerek”, like a parakeet, the end a short descending “churr”; also a soft “kokh-kokh-kokh”, a short, high-pitched “kwit”, and clicking sounds in courtship display.

Range and status Indian subcontinent in Pakistan (Indus Valley fron NE Punjab to SE Sind), India (in north from Kumaon and the Bhutan duars to W Bengal and in S Nepal (Karnali Valley) south to Eastern Ghats, Deccan and the Nilgiris), and in dry parts of Sri Lanka. Resident. Occur from lowlands to foothills to the level where tea and rubber plantations have replaced the indigenous scrub and forest. The species range expanded during first half of 20th century into drier areas of Sind along irrigation canals (Inglis et al. 1920, Ali 1953, Ali and Ripley 1969, Roberts 1991, Zacharias and Gaston 1993, Grimmett et al. 1999). Uncommon and local.

thorn and grass jungle of the terai, dry stony hillsides and ravines, semi-desert; plains and foothills to 1000 m, locally as high as 2100 m (Jerdon 1862, Legge 1880, Baker 1934, Ali 1953, Rand and Fleming 1957, Ali and Ripley 1969, Grimmett et al. 1999). Occur in seasonally dry stream beds with scrub jungle and thorny acacia in the Sind region. Terrestrial, feed on the ground, digging with the bill for insects and mollusks, walk under thickets like a coucal, run on the ground with head and tail extended horizontally, stalk through thickets and occasionally fly from tree to tree (Inglis 1903, Roberts 1991, Robertson and Jackson 1992, Ali 1996). Secretive, and when seen, often in pairs.

Food Large insects, mainly grasshoppers, also mantids, caterpillars, termites, lizards, mice, fledged birds, snails, berries and fruits (Jerdon 1862, Legge 1880, Ali and Ripley 1969, Roberts 1991).

Displays and breeding behavior In courtship display, one bird bows and bobs with tail cocked and spread while in front of its mate, bowing the head to the ground then holding it skyward with the bill held open, for a minute or longer (Gill 1923, Ali and Whistler 1937).

Habitat and general habits Dry deciduous secondary forest, scrub and bush, acacias, undergrowth of lantana and thick grass,

Breeding In Pakistan from May to September, in northern India from April to July and August, in southern India (Kerala) in March and May, and in Bengal from May to August (Hume 1873, Hume and Oates 1890, Inglis 1903, Baker 1934, Ali and Whistler 1937, Ali 1953, Gaston 1981, Roberts 1991). The nest is a loose cup or saucer of twigs, lined with dry or green leaves, built near the ground to 4 m above ground, in thickets, vines or scrub, often in a thorn tree, sapling or the fork of a euphorbia. Eggs are chalky white, with a yellowish film when fresh, 36 ⫻ 26 mm, clutch 2(3) (Hume and Oates 1890, Baker 1934). Both sexes incubate. Incubation and nestling periods are unknown.

Red-billed Malkoha Zanclostomus javanicus 293

Genus Zanclostomus Swainson, 1837 Zanclostomus Swainson, 1837. Classification of Birds, 2, p. 323. Type, by monotypy, Phoenicophaeus javanicus Horsfield 1821. The genus name refers to the

sickle-shaped bill (Gr. zanclon, sickle; stoma, mouth). The genus is characterized by the feathered face with a small unfeathered patch. One species.

Red-billed Malkoha Zanclostomus javanicus (Horsfield, 1821) Phoenicophaus Javanicus Horsfield, 1821, Transactions of the Linnean Society of London, 13 (1), p. 178. ( Java) Polytypic. Two subspecies. Zanclostomus javanicus javanicus (Horsfield 1821), Zanclostomus javanicus pallidus Robinson and Kloss 1921.

Description ADULT: Sexes alike, above head gray, back and wing glossy bluish green, rump slate gray, tail glossy blue gray with white tips, tail feathers finely barred (as in male Raffles’s Malkoha Rhinortha chlorophaea) when seen at close range, face gray with rufous lores; underparts, throat and breast light rufous, lower breast light rufous and gray, belly and under tail coverts rufous, under wing coverts gray; bare skin around eye gray or blue, the patch smaller than in other malkohas, iris white (or light brown, deep brown or red), bill red, base and culmen blackish, legs slate gray. JUVENILE: Like adult, but rectrices narrow (T1 maximum width 30 mm in juvenile, 40 mm in adult), white tip of tail shorter (T1, juvenile 8 mm, vs adult 18 mm). In fresh plumage, primary coverts with rufous wash and pale edges, and bill red with black tip. NESTLING: Undescribed. SOURCES: AMNH, BMNH, CM, FMNH, MZB, RMNH, ROM, SMF, UMMZ, USNM, ZRC.

Zanclostomus javanicus pallidus Robinson and Kloss 1921; lower breast pale gray; peninsular Burma and Thailand, Malay Peninsula, Sumatra, Natuna Is, Borneo.

Measurements and weights Z. j. javanicus, wing, M (n ⫽ 6) 135–147 (134.7 ⫾ 7.7), F (n ⫽ 3) 135–143 (138.3) (AMNH, MZB); Z. j. pallidus, wing, wing, M (n ⫽ 16) 136-160 (144.3 ⫾ 6.2), F (n ⫽ 15) 140–154 (147.3 ⫾ 4.9); tail, M 225–275 (249.4 ⫾ 13.8), F 233–288 (265.6 ⫾ 17.1); bill, M 31–37 (33.8 ⫾ 2.1), F 31.5– 37 (34.1 ⫾ 1.6); tarsus, M 26.5–31 (30.8 ⫾ 2.07), F 29–34 (30.6 ⫾ 1.9) (AMNH, ZRC). Weight, Z. j. javanicus, M (n ⫽ 1) 90 (MZB); Z. j. pallidus, M (n ⫽ 1) 98.0, F (n ⫽ 1) 97.0 (Thompson 1966). Wing formula, P6 ⬎ 5 ⬎ 4 ⫽ 7 ⬎ 3 ⬎ 2 ⬎ 8 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 42 cm. Malkoha with plumage gray above and light rufous below with a gray breast, a very long tail and a red bill.

Voice A frog–like doubled “turk turk urk-urk”, sometimes ending in a rattled flurry of “kuk kuk” like a squirrel; a hoarse “tok” or “turk-urk”, a whining “taup”, and a clear whistled “who-oo” repeated every 10 sec (Lekagul and Round 1991,Wilkinson 1991, MacKinnon and Phillipps 1993, Robson 2000a, Sheldon et al. 2001, NSA).

Subspecies

Range and status

Zanclostomus javanicus javanicus (Horsfield, 1821); lower breast dark gray; Java;

Southern Burma (Tenasserim), southern Thailand, Malay Peninsula, Sumatra, Natuna Islands, Java and

294 Chestnut-breasted Malkoha Phaenicophaeus curvirostris regrowth and hill slopes. In Sabah most records are below 500 m; a bird at 1300 m at Rinangisan may have dispersed during the El Niño drought of 1982 (Sheldon et al. 2001).The birds occur in pairs, and in feeding sometimes they associate with coucals Centropus spp. (Horsfield, in Kuroda 1936).

Food

Borneo (Büttikofer 1900, Smythies 1940, 1981, van Marle and Voous 1988, Lekagul and Round 1991, MacKinnon and Phillipps 1993, Wells 1999, Robson 2000a). Resident.

Habitat and general habits High trees in drier forest, evergreen and semievergreen lowland and lower montane forest, in forest edge and midstory, also in logged forest, secondary scrub, and cocoa, Albizia and Eucalyptus plantations; in lowlands and hilly areas (King and Dickinson 1975, Wilkinson et al. 1991,Wells 1999,Sheldon et al. 2001). In the Malay Peninsula the birds are regular and uncommon to common from the plains to 1200 m, in elfin forest on high steep ridges, in scrub, and in forest

Large insects, including cicadas, walking sticks (phasmids), caterpillars (Sphingidae) (up to 8 cm and 5 g), insect pupae (Pieridae), grasshoppers (Acrididae, Tettigoniidae), mantids, cockroaches, katydids, beetles (Scarabaeidae), bugs and winged termites; spiders and crustacea (Sody 1989,Wells 1999, Sheldon et al. 2001; USNM).

Breeding In southeast Asia they breed from May to July (Robson 2000a), in the Malay Peninsula in June (Wells 1999), in southern Borneo in February and April (Holmes and Burton 1987, Holmes 1997, Smythies 1999). The nest is a flimsy platform of twigs (Chasen 1939, Hellebrekers and Hoogerwerf 1967). Eggs are chalky white, 29 ⫻ 23 mm (Hoogerwerf 1949, Schönwetter 1964). Clutch size is 2 (Robson 2000a). Incubation and nestling periods are unknown.

Genus Phaenicophaeus Stephens, 1815 Phaenicophaeus Stephens, 1815, Shaw’s General Zoology, 9, pt. 1, p. 58. Malkohas are large, slender cuckoos with a brightly colored bill, a brightly colored bare face, and a long tail. Type, by subsequent designation, Cuculus pyrrhocephalus Pennant 1769 (Peters 1940, p. 56). The genus name refers to the crimson appearance (Gr. phoinikophaes) in Red-faced Malkoha P. pyrrhocephalus.The name is often spelled

“Phoenicophaeus” or “Phoenicophaës” (as in Cabanis and Heine 1843), but Stephens 1815 spelled it “Phaenicophaeus” and the “-oe” variant derives from later faulty typesetting of diphthongs “ae” and “oe”, examples in ICZN (1985). Includes two other genera of Peters (1940), Rhopodytes and Rhamphococcyx (part), all with brightly colored bare face, no crest, and no sexual dimorphism in plumage. Six species.

Chestnut-breasted Malkoha Phaenicophaeus curvirostris (Shaw, 1810) Cuculus curvirostris Shaw 1810, Naturalists’ Miscellany, 21, p. 905. [western Java]

Polytypic. Six subspecies. Phaenicophaeus curvirostris curvirostris (Shaw, 1810); Phaenicophaeus

Chestnut-breasted Malkoha Phaenicophaeus curvirostris 295 curvirostris oeneicaudus Verreaux and Verreaux, 1855; Phaenicophaeus curvirostris harringtoni Sharpe, 1877; Phaenicophaeus curvirostris microrhinus Berlepsch, 1895; Phaenicophaeus curvirostris singularis Parrot, 1907; Phaenicophaeus curvirostris deningeri Stresemann, 1913. Other names: Rhamphococcyx curvirostris (Shaw 1810).

Description ADULT: Sexes similar, alike, head gray, back and wing dark glossy green (blue in worn plumage), tail greenish black and rufous with no white tips, T1 dark glossy green and the distal part dark rufous,T2 to T5 increasingly more extensively rufous with outer T5 rufous except at the tip, the tail strongly graduated, green above and rufous below, the face with cheeks [rufous and [gray, in male the gray streak under the eye constricted and ending at the bill, in female the gray streak broad and continuous with gray chin; underparts, in male the chin rufous, in female the chin gray; underparts, throat and breast rufous, belly and under tail coverts dark rufous to black, under wing coverts dark rufous to black; bare skin around eye rugous (not smooth) and red, the bare skin extending forward to the nostril, the iris light blue with an outer ring of paler blue in male, iris yellow in female, bill arched, deep and swollen in shape and in some subspecies with a groove running from nostril to edge of the bill, the bill light green above, red or black below, feet dark gray [photos, Ong Kiem Sian]. JUVENILE: Plumage like adult, tail feathers narrower (T1, juvenile, 35–36 mm; adult, 46–47 mm), dark gray central rectrices less green than in adult; underparts, throat and breast dark purplish rufous, belly and under tail coverts dull black; less bare facial skin than adult, iris dark brown, later changing to reddish brown or white (Hume and Davison 1878), bill dark gray, not arched and swollen and lacks the groove. NESTLING: Skin gray with prominent hair-like down, bill grayish-blue. By day 10 the nestling is well feathered, orbital skin is dull dusky red (Madoc 1956a).

SOURCES: AMNH, BMNH, CM, DMNH, FMNH, FU, MZB, RMNH, ROM, UMMZ, USNM, ZRC, ZSM.

Subspecies Phaenicophaeus curvirostris singularis (Parrot, 1907); crown slate gray, back and wings bronzy green, narrow gray line under eye, slightly broader in female, cheeks, throat and breast rufous brown, belly and under tail coverts blackish; bare face finely outlined with small feathers of white and black, nostril round to oval, wider than deep, the bill pale green with red below nostril and red on lower mandible, bill has either no groove or a vertical groove from nostril to cutting edge; southern Burma, southern Thailand, Malay Peninsula, Anamba Islands (Pulo Siantan) and Sumatra; Phaenicophaeus curvirostris oeneicaudusVerreaux and Verreaux, 1855; plumage darker, the crown and nape black, back and wings dark glossy green, tail with the central rectrices all green lacking any rufous tip or brown and appearing dark green both above and below, face dark maroon and slate, throat to belly dark vinous chestnut, under tail coverts black, bill as in P. c. singularis; western Sumatra islands; Phaenicophaeus curvirostris curvirostris (Shaw, 1810); crown greenish gray, back dark bronzy green, gray streak below the eye straighter and broader in female than in male, underparts rufous, paler than in P. c. singularis, the throat paler than the breast and belly; the bill green above and black below, nostril slitlike to oval, groove runs from nostril forward along base of bill; west and central Java; Phaenicophaeus curvirostris deningeri Stresemann, 1913; underparts rufous, slightly paler than P. c. curvirostris; bill as in P. c. curvirostris; east Java and Bali; Phaenicophaeus curvirostris microrhinus Berlepsch, 1895; crown gray, back bronzy green, tail with broader rufous tip than in P. c. curvirostris, the face dark gray; underparts reddish brown, darker than P. c. singularis; bill green with red below nostril extending forward along cutting edge and red on lower mandible, nostril oval, wider than deep, no groove; Borneo and Bangka; Phaenicophaeus curvirostris harringtoni (Sharpe, 1877); above as in P. c. microrhinus, breast bright rufous, narrow and inconspicuous streak of gray

296 Chestnut-breasted Malkoha Phaenicophaeus curvirostris below eye slightly broader in female than in male, gray throat of female reduced compared with other subspecies and the sexes not readily distinguished by plumage, belly and under tail coverts rufous, tail with T1 having the terminal third vinaceous and the outer two T4 and T5 all vinaceous (more extensive vinaceous than in other forms, both adult and juvenile); the nostril a vertical slit in the feathered base of bill, no groove on the bill; bill green above and red below; western Philippine Islands (Balabac, Palawan, Dumaran I and Calamian Is). Description and taxonomic notes: although confused in earlier literature (Hume and Davison 1878, Büttikofer 1900, Robinson and Kloss 1923b, Robson 2000a, 2002), the iris color of adults differs between the sexes. Specimen labels having both sex and iris color consistently indicate the iris blue (light blue, white, gray, and pantj (⫽ gray-blue, pantj is an old Indonesian term for a cooking pot) in males, and yellow (including cream, orange and red) in females; the few exceptions are mis-sexed (as shown by chin color), or colors mis-translated from one language on the field label to another on a museum label (German, Dutch, English, Sundanese, Malay, Bahasa Indonesia, cf.Wallace 1869 Appendix), or immature (iris brown or dark brown). For P. c. oenicaudatus, male iris color is reported as light blue and female as dark brown; for P. c. harringtoni iris color is unknown. The subspecies name of the Malay and Sumatra form is P. c. singularis Parrot 1907 rather than erythrognathus Bonaparte 1850, for nomenclatural reasons (Deignan 1952).

Measurements and weights P. c. singularis: Malay Peninsula and Sumatra, wing, M (n ⫽ 9) 160-180 (169.1 ⫾ 7.9), F (n ⫽ 11) 166– 182 (171.5 ⫾ 5.5); tail, M 240–274 (256.7 ⫾ 11.4), F 245–290 (264.3 ⫾ 12.6); bill, M 37–42.5 (40.4 ⫾ 1.9), F 37–42.5 (39.9 ⫾ 1.5); tarsus, M 36.5–44.5 (39.1 ⫾ 2.3), F 37–41 (38.6 ⫾ 1.9) (AMNH, BMNH); P. c. oeneicaudus, Mentawai Is, wing, M (n ⫽ 5) 159–174 (165.4), F (n ⫽ 6) 163–175 (168.7 ⫾ 4.2) (MZB); P. c. curvirostris, West Java, wing, M (n ⫽ 6) 166–177 (171.2 ⫾ 4.5), F (n ⫽ 6) 170–174 (171.7 ⫾ 1.9) (MZB);

P. c. microrhinus, Kalimantan, Borneo, wing, M (n ⫽ 4) 154–165 (159.0), F (n ⫽ 7) 156–169 (162.4 ⫾ 4.7) (MZB); P. c. harringtoni, Palawan, wing, M (n ⫽ 8) 171– 181 (175.5 ⫾ 2.8), F (n ⫽ 6) 167-180.5 (175.4 ⫾ 4.2) (DMNH, USNM). Weight, P. c. singularis, Malay Peninsula: U (n ⫽ 2) 188, 190 (Wells 1999); P. c. curvirostris, west Java: M (n ⫽ 3) 145–160 (152.0), F (n ⫽ 2) 163–165 (162) (MZB); P. c. microrhinus Kalimantan, Borneo: M (n ⫽ 4) 134–155 (142.3), F (n ⫽ 7) (141.9) (MZB); P. c. harringtoni, Palawan: M (n ⫽ 6) 111–177.1 (152.3), F (n ⫽ 3) 137.5–148.6 (144.2) (FMNH, USNM,YPM). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 42–49 cm. Large malkoha with gray head, short wing and long tail with a dark tip, body dark glossy green above and chestnut throat and breast, bare red face, bill strongly arched, light green above and dark below, and red below in part of the species range.Wallace (1869) described the birds in Malacca as “green and brown cuckoos with velvety faces and green beaks”.

Voice Call, a loud “wee-oo”, lower on the second note, perhaps the same call described as a slow, low “kukkuk” like a clucking chicken, and a fast “kok-kokkok” when disturbed or in flight and given 2 notes per sec, a mewing call “miaou” with two feeding birds calling to each other; and a single gentle “konk” when approaching a nest (Madoc 1956a, MacKinnon and Phillipps 1993, Robson 2000a, Sheldon et al. 2001).

Range and status Southeast Asia in southern Burma (Tenasserim), southwestern Thailand, the Malay Peninsula and Anamba Islands through Sumatra, Borneo, Java, Bali and neighboring islands, and the western Philippines (Palawan group) (Hume and Davison 1878, Oberholser 1917, Chasen 1939, Stresemann

Chestnut-breasted Malkoha Phaenicophaeus curvirostris 297

Food Insects, caterpillars (hairy and hairless, Lasiocampidae), grasshoppers (Acrididae), cicadas, beetles, cockroaches, stick insects (to 150 mm long), locusts, mantids; small crabs, frogs, nestling birds, lizards (to 280 mm long), small snakes and rats, and leaves (Chasen 1939, Madoc 1956a, Smythies 1981, Gonzales and Rees 1988, Sody 1989, Strange and Jeyarajasingam 1993, Wells 1999, Sheldon et al. 2001; AMNH).

Breeding 1913b, du Pont 1971, King and Dickinson 1975, Smythies 1999, Mees 1986, 1996, Holmes and Burton 1987, van Marle and Voous 1988, Dickinson et al. 1991, Lekagul and Round 1991, Davison and Fook 1995, 1996,Wells 1999, Robson 2002). Resident. Locally common in lowland forests in Thailand, Borneo and the Philippines. The forests where they live are being cut and destroyed.

Habitat and general habits Thickets in lowland forest, forest edge, in dense areas in wind-throw gaps and selectively logged forests, second growth, peat swamp forest, mangroves, creepers, vines, old plantations of cocoa, citrus and rubber, and in gardens. They occur on the mainland and on islands from lowlands to about 1000 m (Smythies 1981, van Marle and Voous 1988, van Balen and Prentice 1997, Wells 1999, Kemp 2000). They live in primary and secondary forest and into mangroves on Langkawi I, Malay Peninsula, and in the Mentawai Islands, Sumatra. In forest canopy 30–50 m above ground (Thompson 1966) they perch motionless, watching and waiting, then pounce, sometimes descending to the ground to pick up prey. In a tree they move and hop and swing their tails like a squirrel. Usually solitary, they occasionally feed in mixed-species flocks with other malkohas, drongos, woodshrikes and magpies (Madoc 1956a, Gonzales and Rees 1988, Wells 1999, Sheldon et al. 2001).

In southeast Asia from January to September (Robson 2000a), in Malay Peninsula with eggs from January to March, nestlings in June and July, and a fledgling begging from parent on 26 September (Madoc 1956a, Wells 1999), in Sumatra in January (eggs) and in Java nestlings and fledged young from January to April, in July, August and November (Baker 1934, Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967, van Marle and Voous 1988; RMNH), in Borneo in February and from August to December (Smythies 1957, 1999). Nest is a thick platform of dead twigs (often casuarinas or rambutan), sometimes with the leaves attached, with a small and shallow cup for eggs lined with leaves; the nest outside 35 cm in diameter and 13 cm deep, the cup 11 cm across and 5 cm deep. The nest is built in the fork of a tall bush or in a tree as high as 2.5–10 m above ground. Eggs are chalky white, becoming stained by dirt, size 34 ⫻ 28 mm the clutch 2–3 (Baker 1934, Chasen 1939, Hoogerwerf 1949, Madoc 1956a, Sheldon et al. 2001). Incubation period is 13 days or longer. Both sexes incubate, care for the brood and clean away the nest debris. Parental feeding trips to the nest are infrequent, once an hour, when they bring the nestlings large items, mainly grasshoppers. The parent retains hold on one end of a large insect in its bill for several minutes after it places the other end into the nestling’s bill, and it takes back the insect when the young does not swallow it. Between feeds the young give soft bubbling calls; when fed they give groans”. Nestling period is at least 11 days (Madoc 1956a,Wells 1999).

298 Chestnut-bellied Malkoha Phaenicophaeus sumatranus

Chestnut-bellied Malkoha Phaenicophaeus sumatranus (Raffles, 1822) Cuculus Sumatranus Raffles, 1822, Transactions of the Linnean Society of London 13, p. 287. (Sumatra and the adjacent islands) Other common names: Rufous-bellied Malkoha. Other names: Rhopodytes sumatranus (Raffles]. Monotypic.

Description ADULT: Sexes alike, head gray, back dark glossy green, wing glossy greenish blue, tail glossy bluish gray with white tips, face with a border of black around the brightly colored bare skin; underparts, throat gray and breast dark blue gray, belly and under tail coverts rufous chestnut; velvety orangered skin around eye, iris pale blue to white, bill pale green, feet gray to gray green. JUVENILE: Like adult, rectrices narrower and less squared at the tip, tail with white tips shorter and narrow iris brown. NESTLING: Skin purplish black, with short yellowish hair-like down tips of growing contour feathers; bare skin on face grayish-pink in first week, changing to dull orange-red by day 14 iris brown, mouth lining red, palate with raised white rugose papillae (one on each side of palate, two smaller papillae in a lateral row behind each large papilla, and tongue red with a white papilla on the surface (Ong Keim Sian, photos).

Weight, Kalimantan, M (n ⫽ 5) 82–95 (86.0). F (n ⫽ 3) 90–105 (98.3) (MZB). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 8 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

History Meyer de Schauensee (1940) noted a stouter bill in birds from the Batu Islands, West Sumatra; Ripley (1942) described these birds as P. s. rodolphi, larger than elsewhere. Mees (1986) found no geographic variation in size across this range, and there is no obvious variation in plumage color. Birds in Borneo were described as a smaller subspecies P. s. minor (Riley 1938) but are not smaller.

Field characters Overall length 40 cm. Dark greenish cuckoo with gray head, a long tail and white tip, gray to dark rufous underparts, a green bill, and orange skin around the pale blue eye. The white in the tail tip is shorter than in Black-bellied Malkoha P. diardi and Green-billed Malkoha P. tristis.

Voice It makes a knocking sound “kokokokokoko.” at about 12 “ko” notes per sec and a pitch of 1.4 kHz, a simple “chak”, and a thin high-pitched “mew” descending from 2.4 to 1.8 kHz and lasting about 0.3 sec (Robson 2000a, Supari 2003).

Range and status SOURCES: AMNH, BMNH, CM, FMNH, MVZ, MZB, RMNH, ROM, SMF, UMMZ, USNM, ZRC.

Measurements and weights Malay Peninsula: Wing, M (n ⫽ 5) 134–145 (139.4 ⫾ 4.3), F (n ⫽ 5) 135–142 (138.6 ⫾ 2.7); tail, M 220–244 (232.4 ⫾ 11.1), F 219–252 (233.8 ⫾ 18.6); bill, M 33–37 (34.7 ⫾ 1.5), F 29–35 (33.4 ⫾ 2.5); tarsus, M 28–31 (29.9 ⫾ 1.2), F 29–34 (31.1 ⫾ 2.0); Borneo: Wing, M (n ⫽ 6) 139–144 (141.8 ⫾ 2.5), F (n ⫽ 5) 130–146 (136.6 ⫾ 6.4) (AMNH).

Southeast Asia in western Thailand, peninsular Burma (Tenasserim), and the Malay Peninsula, and in Sumatra and Borneo (Hume and Davison 1878, Büttikofer 1900, Smythies 1940, Lekagul and Round 1991, MacKinnon and Phillipps 1993,Wells 1999, Robson 2000a). Resident. Conservation status, near-threatened (BirdLife Inaternational 2001).

Habitat and general habits Lowland forest and forest edge, second growth, peat swamp forest, mangroves, rubber and durian plantations, in dense crowns of trees; lowlands to 1200 m. They live in forest, overgrown plantations and man-

Red-faced Malkoha Phaenicophaeus pyrrhocephalus 299 1992).Their behavior is like that of other malkohas, secretive, creeping through the canopy, and moving in short flights from tree to tree.

Food Insects, including locusts, cicadas, mantids, stick insects (to 20 cm), crickets, grasshoppers, katydids, large caterpillars; lizards, small fruits and seeds (Raffles 1822, Smythies 1981, Hails 1987, Wells 1999, Sheldon et al. 2001).

Breeding groves in Malay Peninsula (Medway and Wells 1976), and are absent on hills over most of the region. In Singapore, where these are now the only malkoha, seen in the forested water catchment nature reserves (Ong Kiem Sian,Wang Luan Keng). They live in overgrown rubber plantations and mangroves in Sumatra (van Marle and Voous 1988) and in coastal areas, riverine forests and logged peat swamp forests in Borneo (van Balen and Prentice 1997, Sheldon et al. 2001). Quiet, they feed as they creep through saplings and thickets (Chasen 1939, Duckworth et al. 1996, Smythies 1999), and they feed in mixed-species flocks (Wilkinson et al. 1991). They are sometimes kept in aviculture (Raethel

In the Malay Peninsula they nest from January to July (Wells 1999) and a male had a brood patch in October (USNM 486813), in Borneo young were found in July and females were in laying condition in October and in December (Smythies 1957, 1999, Nash and Nash 1988, Sheldon et al. 2001). Nest is a flat platform of twigs and bits of vine stem, lined with green leaves, built in a low shrub or high in a tree. Eggs are white, 27.5 ⫻ 24 to 30 ⫻ 23 mm, clutch 2 (Coomans de Ruiter 1947, Smythies 1981, Wells 1999).The male takes part in incubation and parental care. Nestlings beg by stretching the neck upwards and wings out at the side. Incubation period unknown, nestling period 14 days. In Singapore a pair nests repeatedly through the year (Ong Kiem Sian).

Red-faced Malkoha Phaenicophaeus pyrrhocephalus (Pennant, 1769) Cuculus pyrrhocephalus Pennant, 1769, Indian Zoology, p. 6, plate 6. (Ceylon) Monotypic. Other common names: Crimson-faced Malkoha.

JUVENILE: Smaller red face patch, crown streaked brown, upperparts less glossy than in adult, throat darker, breast black streaked with white.

Description

NESTLING: Undescribed.

ADULT: Sexes alike, head and back black glossed green and blue, crown and neck finely streaked white, wing black, tail black with very broad white tips; underparts, chin and lower face whitish, throat and upper breast black, lower breast and belly to under tail coverts white; bare velvety red skin on face, iris brown (males) or white ( females), bill stout, curved, light green above and bluish green below, feet slaty gray.

SOURCES: AMNH, BMNH, FMNH, ROM, SMF, USNM,YPM.

Measurements Wing, M (n ⫽ 6) 148–169 (154.7 ⫾ 6.7), F (n ⫽ 6) 151–169 (157.0 ⫾ 8.1); tail, M 279–318 (290.2 ⫾ 14.5), F 272–300 (248.7 ⫾ 10.9); bill, M 33–35 (34.1 ⫾ 0.8), F 32–35.4 (33.9 ⫾ 1.4); tarsus,

300 Red-faced Malkoha Phaenicophaeus pyrrhocephalus M 29.2–34.2 (31.8 ⫾ 1.7), F 31.0–34.8 (33.1 ⫾ 1.4) (AMNH, FMNH, UMMZ, USNM). Wing formula, P6 ⬎ 5 ⬎ 7 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 8 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 47 cm. Large malkoha with pied plumage, red face, green bill, and black and white underparts, and very long tail. In tall trees, the field marks to note are the extensive white ventral plumage and long tail, appearing white from below. The red face and green bill also are distinctive.

Voice Usually silent, it gives a harsh “snarr, snarr” or a soft cricket-like “kree-kree”, a low whining “kra”, a “kok”, and short single yelping whistles (Fleming 1977, NSA).

Range and status Sri Lanka. Resident. Locally common, generally scarce and limited to small local populations in the wet zone of southern Sri Lanka. In the nineteenth century it occurred in most forests and jungles of the low country especially in the eastern districts (Legge 1880), but its range is now reduced and highly fragmented. In recent years since 1980 it has been observed northeast of the central highlands in Wasgomuwa and Heen Ganga, near the southeast coast at Lahagula Kitulana Sanctuary, Kumbukkan Oya and Yala, and in southwestern Sri Lanka at Kitulgala, Peak Wilderness Sanctuary and Uda Walawe National Park to the southwest including Walankanda Forest Reserve and Yagirala.

populations have declined with the loss of forest habitat to logging, firewood gathering, clearance of forest for permanent cultivation, mining, fire and urbanization. During a survey of more than 200 forest sites in Sri Lanka in the early 1990s, the species was recorded in only three of these forests (BirdLife International 2001). In southern India there are questioned sight records in Kerala, in Tamil Nadu, and Coorg (Kodagu) in S Karnataka, where many birdwatchers have visited but none have seen these birds (Biddulph 1956, Hoffmann 1989a, 1996, Kotagama and Fernando 1994, BirdLife International 2001). Population size a few thousand Globally threatened, vulnerable (BirdLife International 2001).

Habitat and general habits The species lives in tall forest with dense tangled undergrowth in the wet zone of Sri Lanka. They also occur in riverine forests in the dry zone of eastern Sri Lanka. Most records are in the lowlands and records above 920 m are rare, though the birds have been seen to 1540 m at Haputale but probably no longer occur at high altitudes (Henry 1971, BirdLife International 2001). The cuckoos live in the canopy of tall trees, where they hop from branch to branch through tangled twigs, creepers and foliage, and they flutter and glide from tree to tree. Usually they are seen in trees as single birds or pairs; they also occur in flocks of 6–7 birds (Legge 1880), larger than a family group; and they associate with other birds in mixed-species flocks (Fleming 1977, BirdLife International 2001).

Food Fruits and berries of forest trees, and insects, including caterpillars plucked from the vegetation (Legge 1880, Fleming 1977, Fuller and Erritzøe 1997, BirdLife International 2001).

Breeding behavior Little known. A breeding pair was observed together, the male following the female and giving a soft call, “kree-kree-kree”.

Breeding From January to May; gonads are enlarged through August and September.The nest is a shallow saucer

Blue-faced Malkoha Phaenicophaeus viridirostris 301 of interlocking branches, roots and twigs, lined with leaves and leaf midribs. It is built in a shrub near ground in dense undergrowth or well concealed 9–18 m above the ground. Eggs are chalky white, 36 ⫻ 27 mm, clutch is 2–3. Both sexes build

the nest and feed the young. The incubation and nestling periods are unknown (Legge 1880, Lewis 1898, Baker 1934, Whistler 1944, Ali and Ripley 1969, Fuller and Erritzøe 1997, BirdLife International 2001).

Blue-faced Malkoha Phaenicophaeus viridirostris ( Jerdon, 1840) Zanclostomus viridirostris, Jerdon, 1840, Madras Journal of Literature and Science, 11, p. 223. (bottom of the Coonoor Pass [India]) Monotypic. Other names: Rhopodytes viridirostris ( Jerdon, 1840). Other common names: Lesser Green-billed Malkoha, Small Greenbilled Malkoha.

Description ADULT: Sexes alike, head slate gray, back greenish gray, wing glossy blue green, tail long, graduated, feathers broad (T1, 25 mm), black with green gloss and white tips (T1, 32 mm;T2, 38 mm,T3, 34 mm; T4, 35 mm; T5, 30 mm); underparts, throat and breast dark gray with fine whitish streaks, feather hackles bifurcate forked at tip and barbs clump in a point, looking black and bristly, belly gray and light rufous buff, under tail coverts dark gray with light rufous tips; colorful bare pale blue skin around eye, iris narrow white outer ring, red inner ring, bill pale, yellowish to horn green, feet slaty blue. JUVENILE: Upperparts less glossed green, underparts pale gray with no rufous or buff, no bifurcate hackles, outer primary softer and broader than in adult where sickle-shaped, tail feathers are narrower (T1 is 14 mm wide) and with shorter white tip (T1, 5 mm; T2, 7 mm; T3, 22 mm; T4, 26 mm; T5, 14 mm); underparts gray with no rufous or buff. NESTLING: Skin blackish, with long, white hairlike down.

Measurements and weights Wing, M (n ⫽ 13) 132–141 (135.0 ⫾ 3.7), F (n ⫽ 4) 127–138 (131.0); tail, M 220–247 (233.6 ⫾ 10.3), F 222–229 (224.3); bill, M 28.0–31.7 (30.2 ⫾ 1.2), F 29.0–31.0 (30.3); tarsus, M 24–34 (30.6 ⫾ 2.7), F 31–33.6 (32.2) (AMNH, FMNH, UMMZ,YPM). Weight, M (n ⫽ 2) 77–90 (83.5) (USNM,YPM). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 8 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 39 cm. Greenish-gray cuckoo with dark throat and breast and paler belly and vent (the reverse of Green-billed Malkoha P. tristis), bill light greenish, bright blue skin around eye, and long tail with a white tip, the tips more broadly white than in P. tristis. Blue-faced Malkoha live in drier habitats than Green-billed Malkoha P. tristis, and their ranges barely overlap.

Voice A low croak, “kraa” (Ali and Ripley 1969).

Range and status Southern India, mainly in Kerala north to western Karnataka near Goa, Tamil Nadu in the Eastern Ghats and north to the southern Orissa coastal plain, Gujarat; also Sri Lanka (Ali and Whistler 1937, Ali 1953, Ali and Ripley 1969, Abdulali 1971, Robertson and Jackson 1992, Kotagama and Fernando 1994, Grimmett et al. 1999). Resident. Inconspicuous. Widespread, not abundant, most common in dry zone of low country in Sri Lanka. Usually seen as single birds or pairs.

Habitat and general habits SOURCES: AMNH, BMNH, FMNH, MSNG, ROM, SMF, UMMZ, USNM.

Secondary woodland, deciduous thickets, bamboo, thorn scrub and bush with Euphorbia and lantana,

302 Black-bellied Malkoha Phaenicophaeus diardi their short, rounded wings (Lewis 1898, Henry 1971).

Food Insects, caterpillars, grasshoppers, mantids, cicadas, beetles; lizards; fruit (Legge 1880).

Breeding

all dry-country habitats; range is in plains and low foothills to 1000 m. Feed in branches and impenetrable thickets, creep and clamber among branches of low trees, peering on leaves for insects, rarely come to the ground. They move from one thicket to another in a slow, direct and labored flight with

In southern India (Nilgiris) from January to November, mainly March to May (Baker 1934, Ali 1953), in Sri Lanka they breed in all months (Wait 1925, Whistler 1944). Nest is a platform of twigs, lined with green leaves, built near the ground to 2 m above the ground, in thorn scrub, lantana or euphorbia hedge or in a bamboo clump. Eggs are chalky white, 29 ⫻ 25 mm, clutch in India 2(3) (Wait 1925, Baker 1934, Ali and Whistler 1937, Ali 1953, 1996, Ali and Ripley 1969). Incubation and nestling periods are unknown.

Black-bellied Malkoha Phaenicophaeus diardi (Lesson, 1830) Melias Diardi, Lesson, 1830, Traité d’Ornithologie, livre 2, p. 132. (“Java” ⫽ Sumatra) Other names: Rhopodytes diardi (Lesson 1830). Polytypic. Two subspecies. Phaenicophaeus diardi diardi (Lesson, 1830); Phaenicophaeus diardi borneensis (Salvadori, 1874).

adult) and pointed, white tail spots short (T1, 8 mm in juvenile, 18 mm in adult), and no white around the eye, black bristles present; bare skin around eye gray, iris dark brown, bill dark gray.

Description

SOURCES: AMNH, ANSP, BMNH, CM, FMNH, MCZ, MSNG, MZB, RMNH, ROM, UMMZ, USNM.

ADULT: Sexes alike, above dark gray to blackish, wing black with blue-green gloss, tail long, black with blue-green gloss, tipped white, face with a trace of white in the dark feathers around the bare skin, feather shafts of head and anterior body extend into long fine black bristles; underparts gray, belly blackish gray; bare velvety red skin around eye, edge of eyelids black around a bluish-white eye-ring, iris dark brown, bill green with base bluegray, feet dark gray to greenish gray. JUVENILE: Plumage like adult’s but wing sooty, tail feathers narrow (20 mm in juvenile, 30 mm in

NESTLING: Undescribed.

Subspecies Phaenicophaeus diardi diardi (Lesson, 1830); larger, underparts gray; southern Burma (Tenasserim), southern Thailand, Malay Peninsula, Sumatra; Phaenicophaeus diardi borneensis (Salvadori, 1874); smaller, underparts greenish and less gray; Borneo.

Measurements and weights P. d. diardi, Malay Peninsula and Sumatra: Wing, M (n ⫽ 10) 127–138 (131.9 ⫾ 2.8), F (n ⫽ 11)

Black-bellied Malkoha Phaenicophaeus diardi 303 127–137 (130.9 ⫾ 3.0); tail, M 215–240 (227.3 ⫾ 8.4), F 207–241 (222.7 ⫾ 9.3); bill, M 25.4–32 (29.8 ⫾ 1.6), F 28–31 (30.2 ⫾ 1.0); tarsus, M 28.5–30.5 (29.2 ⫾ 0.9), F 28–32 (30.0 ⫾ 1.5) (AMNH, ANSP, BMNH); P. d. borneensis, Borneo: Wing, M (n ⫽ 7) 121– 131 (126.1 ⫾ 3.5), F (n ⫽ 5) 121–136 (126.0), U (n ⫽ 3) 126–128; tail, M 187–198 (193.8 ⫾ 3.9), F 188–217 (197.6) (AMNH, BMNH, RMNH, ZFMK). Weight, P. d. diardi, M (n ⫽ 1) 66, F (n ⫽ 1) 75; P. d. borneensis, M (n ⫽ 1) 62, F (n ⫽ 1) 65 (ANSP, MZB). Wing formula, P7 ⬎ 6 ⬎ 5 ⬎ 8 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 38 cm. Dark greenish cuckoo with gray head, long tail and white tip, dark gray belly, red skin around eye and green bill.The gray of head and breast is darker than in the larger Green-billed Malkoha P. tristis and there is less white around the face.

Voice A loud “pauk”, a gruff, froglike “gwaup” and a twosyllable “gwagaup”; voice also described as “pwewpwew” (Wells 1999, Sheldon et al. 2001).

Range and status Southeast Asia in Laos, Cambodia, southern Burma (Tenasserim) and peninsular Thailand through Malay Peninsula, and in Sumatra and Borneo (Büttikofer

1900, Robinson and Kloss 1923b, Smythies 1940, 1999, King and Dickinson 1975, Mees 1986, Round 1988, van Marle and Voous 1988, Wilkinson et al. 1991, MacKinnon and Phillipps 1993, Duckworth et al. 1996,Wells 1999, Robson 2000a, Sheldon et al. 2001). Resident. Vulnerable to deforestation, they once occurred in Singapore and are now gone, they are rare in Laos and Cambodia, and they are occasional in Sumatra and common to uncommon in lowland forests in Borneo. Conservation status, nearthreatened (BirdLife Inaternational 2001).

Habitat and general habits Forests of primary and second growth, swampy jungle, bamboo, peatswamp, cocoa, Albizia plantations, in dense undergrowth and creepers, less numerous in old rubber plantations and second growth. They occur in lowland plains and foothills to 200 m, sometimes to 1200 m; in Borneo at 120–250 m and to 1000 m, once in Kinabalu NP at 1700 m (Sheldon et al. 2001). Hunt for insects by creeping about in leafy creepers and the outer branches of dense crowns, occur in mixed-species foraging flocks (Medway and Wells 1976,Wilkinson 1991,Wells 1999).

Food Insects, including beetles, ants, mantids, grasshoppers and caterpillars (Chasen 1939, Smythies 1981).

Displays and breeding behavior Both sexes gather nest material (Wells 1999), and the behavior suggests a monogamous pair.

Breeding In Burma in April, in Malay Peninsula from January to April (Baker 1934, Wells 1999), in Borneo in August (Nash and Nash 1985). Nest is a shallow saucer of fine twigs, 18 cm in diameter, 3–4 cm deep, lined with green leaves, built in a tree as high as 5 m above ground in forest edge thicket (Baker 1934, Chasen 1939, Wells 1999). Eggs are chalky white, 30 ⫻ 24.5 mm, clutch size is 2–3 (Baker 1934, Schönwetter 1964). Incubation and nestling periods are unknown.

304 Green-billed Malkoha Phaenicophaeus tristis

Green-billed Malkoha Phaenicophaeus tristis (Lesson, 1830) Melias tristis Lesson, 1830, Traité d’Ornithologie, livre 2, p. 132. (“Sumatra” ⫽ Bengal) Other common names: Greater Green-billed Malkoha, Large Greenbilled Malkoha. Other names: Rhopodytes tristis (Lesson 1830). Polytypic. Two subspecies. Phaenicophaeus tristis tristis (Lesson, 1830), Phaneicophaeus tristis kangeangensis (Vorderman 1893).

Description ADULT: Sexes alike, upperparts dark gray with green gloss, head lighter gray, rump dull slate, wing and tail blackish with green gloss, tail very long with broad white tips, face outlined with white line above extending over the forehead, formed by feathers black at base, white along middle of shaft and black at the tip, lores black, chin whitish, lower margin of face behind the eye with thin black feathered outline, narrow hair-like feathers on forehead, base of bill (where feathers recurved over the nostril) and chin and throat, the black hair-like tips formed by extensions of the black distal barbs; underparts, throat light gray with fine central shaft streaks, breast gray to blackish on lower belly, under tail coverts black; bare red skin around eye, iris brown (or inner ring claret, outer ring white), bill pale green, feet greenish gray. JUVENILE: Like adult but plumage loose-webbed, face gray, head gray with indistinct streaks, no white outline around eye, lores gray, no black hair-like feathers on face, rump dark gray, wing with tips of primaries brownish gray without gloss, rectrices more narrow than in adult and the white tips with about half the area as in the adults, the central tail feathers slightly glossed, narrowly tipped whitish (2 mm), outer tail feathers broadly tipped white, the white markings smaller and less distinct than in adult because tail not as black; bare skin around eye dark brown, iris brown, upper bill slaty horn, lower bill greenish horn, feet pale gray. NESTLING: Undescribed. SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MZB, ROM, UMMZ, USNM, ZRC, ZSM.

Subspecies and Taxonomic Comments Phaenicophaeus tristis tristis (Lesson, 1830; as above; mainland southern and southeast Asia and Sumatra; Phaenicophaeus tristis kangeangensis (Vorderman 1893); above dark gray with olive gloss, tail with large white terminal spots (measured on shaft, T 1, 38 mm, T2,43 mm; T3, 40 mm; T4, 40 mm; T5, 33 mm, each nearly twice the length of P. t. tristis), breast whitish washed yellow; Kangean Is). Oustalet (1899) compared birds from southeast Asia with the series of tristis (Lesson 1830) and elongatus (Müller 1830), and concluded that the southeast Asian birds matched the two birds in the type series of tristis, one from Cochin China and the other from Bengal (Lesson described the birds brought to Paris by the travelers Diard and Duvaucel). Hartert (1910) described Hainan birds as P. t. hainanus, Robinson and Kloss (1919) used this name for their Annam birds, and they reassigned the type locality of Phaenicophaeus tristis to “Bengal”. Bangs and Van Tyne (1931), like Oustalet (1899), noted that neither Hainan birds nor Indochina birds were distinct from the Indian birds. Other forms sometimes recognized include P. t. saliens Mayr, 1938; wing shorter and tail longer; northern Burma, Vietnam (Tonkin) and southern China (Yunnan); P. t. longicaudatus (Blyth, 1842); wing long, tail longer; southern Burma, southern Thailand, Malay Peninsula, Annam and southern Indochina; and P. t. hainanus (Hartert, 1910); wing shorter than saliens, tail shorter, short white tail spots; Hainan. In the measurements below, there is overlap and no great difference in mean size between populations once called tristis and saliens, longicaudatus and hainanus. Birds in the Mishmi hills, northern Assam, are intermediate between tristis and saliens (Ali and Ripley 1948).

Measurements and weights P. t. tristis, Assam: Wing, M (n ⫽ 12) 160–168 (163.3 ⫾ 2.2), F (n ⫽ 11) 151–165 (161.2 ⫾ 4.5); tail, M, 340–408 (374.4 ⫾ 26.8), F, 320–430 (369.8 ⫾ 29.9); bill, M, 30–34 (31.8 ⫾ 1.3), F, 31–32 (31.4 ⫾ 0.5); tarsus, M 33–37 (35.2 ⫾ 1.3), F 33–36

Green-billed Malkoha Phaenicophaeus tristis 305 (34.3 ⫾ 1.0) (UMMZ); Laos: Wing, M (n ⫽ 6) 158–166 (162.3), F (n ⫽ 6) 156–166 (162.3) (Robinson and Kloss 1931); Malay Peninsula and southern southeast Asia: U (n ⫽ 7) wing 148–169 (159.4), tail 361–448 (412.4) (Mayr 1938b), Malay Peninsula: Wing, M (n ⫽ 21) 149–163, F (n ⫽ 10) 153–161; tail M 291–370, F 312–348 (Wells 1999); northern southeast Asia:Wing 152–163 (155.4), tail 364–394 (377.9), once 347 (Mayr 1938b); Hainan: U, wing 148–162 (153.0), tail 319–361 (343.3) (Mayr 1938b); Sumatra, wing M (n ⫽ 3) 145-156 (149.3), F (n ⫽ 5) 151-160 (155.5), tail, M 316-368 (343.3), F (373.3) (MZB); P. t. kangeangensis, wing, M (n ⫽ 6) 160–170 (167.2 ⫾ 3.4), F 160–163 (161.7); tail, M, 332–372 (355⫾13.6), F 342–372 (14.4) (AMNH, MZB). Weight, Nepal, M (n ⫽ 1) 116, F (n ⫽ 1) 114 (ZSM); Kangean Is, M (n ⫽ 2) 120, 128, F (n ⫽ 1) 100 (MZB). Wing formula, P5 ⱖ 4 ⬎ 6 ⬎ 7 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Kangean Is in the East Java Sea (Bangs and Van Tyne 1931, Smythies 1940, Price 1979, van Marle and Voous 1988, Lekagul and Round 1991,Thewlis et al. 1996, Robson et al. 1998, Wells 1999, Evans et al. 2000, Robson 2000a). Common resident. At higher altitudes, a summer visitor in Nepal in Kathmandu (Inskipp and Inskipp 1985), and resident all year in Bhutan at the same altitude as Kathmandu (Ali 1962).

Field characters

Habitat and general habits

Overall length 52–58 cm. Dark green cuckoo with gray head and a pale green bill, red skin bordered white around the eye, underparts grading from pale gray throat to blackish vent (reverse of Blue-faced Malkoha P. viridirostris), and a very long tail with white tips.

Primary forest, second growth, dense thickets, scrub, cultivated areas, coastal scrub and landward fringe of mangroves, paperbark Melaleuca stands, inland evergreen and semi-deciduous forests, overgrown rubber plantations, hill forest and bamboo. They occur from lowlands into hills to 1800 m; in Nepal they are common below 700 m and occur to 1800 m (Jerdon 1862, Ludlow and Kinnear 1937, Smythies 1940, Inskipp and Inskipp 1985, Robson et al. 1998, Grimmett et al. 1999, Wells 1999, Robson 2000a). In the Malay Peninsula a bird of lowland forests in the north, but a hill bird in the south where they live above 850 m on the Main and Larut Ranges. They clamber around like squirrels in creepercovered trees and thickets, where they feed in dense foliage. Tail pattern is seen in flight, spread as the cuckoo approaches a tree or thicket (Strange 2000).

Voice A mellow and nasal well-spaced “oh oh oh oh” repeated on territory, also a croaking frog-like call, “ko, ko, ko,ko” sometimes ending with a gruff flurry “co-co-co-co”, and a cat-like chuckle (Hume and Davison 1878, Smythies 1940, Ali and Ripley 1969, MacKinnon and Phillipps 1993, Robson 2000a).

Range and status India along the lower Himalayas from northern Uttar Pradesh east to Madhya Pradesh and Arunachal Pradesh, Nepal, northeastern and eastern India in the Eastern Ghats, Bengal and in Bangladesh (Biswas 1960, Grimmett et al. 1999), extreme southern China and Hainan (Cheng 1991) through southeast Asia, the Malay Peninsula south to 3° N, Sumatra, and

Food Insects, including large caterpillars and orthoptera; also lizards (Ali and Ripley 1969).

Displays and breeding behavior Birds usually seen in pairs (Inglis et al. 1920).

306 Yellow-billed Malkoha Rhamphococcyx calyorhynchus

Breeding In India they breed from April to August (Oates 1877, Baker 1934, Grimmett et al. 1999), in Assam feathered nestlings are seen in May (UMMZ), in Burma they breed from February to June, or March to September (Oates 1877, Baker 1934, Smythies 1940), in Malay Peninsula from December to March (Wells 1999), in southeast Asia from December to June (Robson 2000). Nest is a shallow platform c. 18 cm across, small for the size of the bird, built of twigs, tendrils and rootlets and lined with green

leaves (green leaves are added through the nesting cycle), built in the middle of a clump of bamboo, creepers or leaves in a tree at height of 3–7 m (Oates 1877, Bingham 1880, Herbert 1924, Baker 1934, Betts 1947,Ali 1962), or perhaps sometimes a closed nest in a tangle of twigs ( Junge and Kooiman 1951). Eggs are chalky white, 34 ⫻ 26 mm, clutch 3(2–4) in India, 2–3 in Burma, 2 in Thailand and Malay Peninsula. Both sexes incubate. Incubation and nestling periods are unknown (Ali 1996, Wells 1999).

Genus Rhamphococcyx Cabanis and Heine, 1863 Rhamphococcyx Cabanis and Heine, 1863, Museum Heineanum, Thiere, 4, 1862–1863, Heft 1, 65. Type, by original designation, Phaenicophaeus calyorhynchus Temminck, 1825. The name refers to the large bill (Gr. rhamphos, the bill; Gr. kokkux, the cuckoo).

One species. Chestnut-breasted Malkoha Phaenicophaeus curvirostris, is often included in the genus, but is excluded here because of the molecular genetic results.

Yellow-billed Malkoha Rhamphococcyx calyorhynchus (Temminck, 1825) Phaenicophaeus calyorhynchus Temminck, 1825, Nouveau recueil de planches coloriées d’oiseaux, livre 59, pl. 349. (Celebes) Other common names: Fiery-billed Malkoha. Polytypic. Three subspecies. Rhamphococcyx calyorhynchus calyorhynchus (Temminck 1825), Rhamphococcyx calyorhynchus meridionalis (Mayer and Wiglesworth 1896), Rhamphococcyx calyorhynchus rufiloris 1903.

elongate and horizontal, no groove in the bill, feet black. JUVENILE: Plumage like adult, but crown rufous gray, rectrices narrow (T1 is 34 mm in width, vs 40 mm in adult), lower belly brownish gray; iris brown, bill blackish, not swollen and arched as in adult, the gonys with a distinct ridge. NESTLING: Undescribed.

Description ADULT: Sexes alike, crown dark gray, back rufous maroon, lower back and rump purplish black, wing dark purple, tail purplish black, face dark gray, with gray or rufous feathers above the bare lores, cheeks rufous; underparts, throat and breast rufous, belly dark gray, underwing coverts black; bare skin around eye red to black, iris red, bill strongly arched, deep and swollen in shape, above yellow with a black band near the tip, bill below the nostril red and lower mandible red, nostril

SOURCES: AMNH, CM, FMNH, MZB, RMNH, ROM, SMTD, UMMZ, USNM, ZMB.

Subspecies Rhamphococcyx calyorhynchus calyorhynchus (Temmick 1825); plumage, crown dark gray, breast rufous chestnut; northern Sulawesi and Togian Is; Rhamphococcyx calyorhynchus meridionalis Meyer and Wiglesworth 1896; plumage, crown ligher gray, breast ochre; southern Sulawesi;

Yellow-billed Malkoha Rhamphococcyx calyorhynchus 307 Rhamphococcyx calyorhynchus rufiloris (Hartert 1903), Buton Is; plumage as in R. c. calyorhynchus, smaller.

Measurements R. c. calyorhynchus, wing, M (n ⫽ 10) 164–178 (175.4 ⫾ 5.3), F (n ⫽ 8) 172–189 (177.8 ⫾ 4.9); tail, M 302–344 (320.9 ⫾ 17.8), F 324–344 (331.3 ⫾ 5.5); bill, M 40.5–45 (42.8 ⫾ 1.3), F 41–45 (42.8 ⫾ 1.5); tarsus, M 34–38 (37.0 ⫾ 1.4), F 34.5–39 (37.1 ⫾ 1.9) (AMNH, MZB, UMMZ); R. c. meridionalis, wing, M (n ⫽ 8) 172–193 (180.4 ⫾ 7.3), F (n ⫽ 6) 168-187 (179.8 ⫾ 8.1) (AMNH, MZB); R. c. rufiloris, wing, M (n ⫽ 2) 165, 174, F (n ⫽ 2) 162, 174, tail, M 290, F 294, 330 (AMNH, MZB). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 3 ⬎ 2 ⫽ 8 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 53 cm. Cuckoo with chestnut back and breast, gray crown and red orbital skin, short wing and long purplish-black tail, and the bill swollen and strongly arched, yellow above and black at tip, and red below.

Voice Calls, a nasal rattle or chatter which accelerates then dies away, and rises and falls in pitch (squirrel-like chatter at rate of 8–10/sec at a pitch between 3 and 4 kHz, lasting 1.5 sec, and a nasal whining note like a branch cracking in the breeze (Watling 1983, tape in NSA; Holmes and Phillipps 1996; Thomas and Thomas 1994; Coates and Bishop 1997). The bird is said to have ten different calls; and local people look upon it as a prophet of the day, conversing with it by imitating the calls and attracting the calling bird (Meyer 1879).

Range and status Sulawesi, Butung and neighboring islands. Resident. Widespread and common. They are seen in Tangkoko Nature Reserve, Rawa Aopa Watumohai NP,Tentolo-Matinan Mts, Lore Lindu NP, Mekonga Mts and the Lomphobattung Massif (White and Bruce 1986, Coates and Bishop 1997). This species

and two in the Philippines are the only malkohas that occur east of Wallace’s line.

Habitat and general habits Primary and tall secondary forest and forest edge, in lowland, hill, and locally lower montane forest, remnant forest patches, scrub, open woodland, and lightly wooded cultivation, also coconut plantations, riverine vegetation near villages; sea level to 1300–1650 m (Heinrich, in Stresemann 1940). Usually solitary. They hop and creep around the canopy like squirrels, peering into dead leaf clusters and epiphytes; and they glide and slide from tree to tree (Meyer 1879). They follow troops of Sulawesi macaque Macaca nigra and booted macaque M. ochreata and catch grasshoppers and other insects flushed out by the monkeys, and they also associate with Bay Coucals Centropus celebensis (Heinrich, in Stresemann 1940, Fooden 1969, White and Bruce 1986, Whitten et al. 1987, Wardill et al. 1998). A feeding association of malkohas with macaques was observed at Tangkoko-Batuangus Reserve, near Manado, NE Sulawesi (White and Bruce 1986); malkohas also occur in coconut plantations where no macaques are found.Two malkohas were said to grasp each other, fight and roll around together on the ground. Common, perching on low shrubs, tail hanging down (Meyer 1879).

Food Insects, including hairy caterpillars, locusts and mantids, and beetles (Meyer 1879, Stresemann 1940).

308 Rough-crested Cuckoo Dasylophus superciliosus

Breeding November to January, with dates estimated from a dependent juvenile observed on 8 January (Rozendaal and Dekker 1989) and a female with a soft-shelled egg on 11 January at Pindepada (Riley 1924). The nest is a flat platform of twigs, like a pigeon (Meyer 1879); this appears to be the only

observation indicating that the malkoha is a nesting species. Egg is white, 36 ⫻ 31 mm (Schönwetter 1964). Incubation and nestling periods are unknown. Although this cuckoo is reported to be a nesting species, a Crimson Sunbird Aethopyga siparaja repeatedly fed a juvenile malkoha (Rozendaal and Dekker 1989).

Genus Dasylophus Swainson, 1837 Dasylophus Swainson, 1837, Classification of Birds, 2, p. 324. Crested malkohas are large, slender cuckoos with a long tail and with a crest of feathers contrasting with the rest of the plumage. Type, by original designation, Phaenicophaeus superciliosus

(Dumont, 1823).The genus name refers to the crest (Gr. dasus, shaggy; lophos, a crest). Includes two genera of Peters (1940), Dasylophus and Lepidogrammus for two Philippines malkohas which are similar as feathered nestlings.Two species.

Rough-crested Cuckoo Dasylophus superciliosus (Dumont, 1823) Phaenicophaus superciliosus Dumont, 1823, Dictionnaire des Sciences Naturelles, éd. Levrault, 28, p. 451. (Philippines) Other common names: Red-crested Malkoha. Polytypic. Two subspecies. Dasylophus superciliosus superciliosus Dumont, 1823; Dasylophus superciliosus cagayanensis (Rand and Rabor, 1967).

Description ADULT: Sexes alike, above black with greenish-blue gloss, head with a bushy, laid-back red eyebrow crest, wing black, tail glossy black with broad white tips; underparts dull black with yellowish-green overtones on throat and belly to under tail coverts, under wing coverts black; bare skin around eye red to yellowish orange, iris yellow, bill pale green with an orange base, feet green in front and yellow behind. JUVENILE: Dull black or black with a brown forehead (the red crest develops shortly after fledging), wing black, tail black with narrow white tips (5 mm), feathers narrower than in adult (T1 width is 20–30 mm in juvenile, 12–15 mm in adult); underparts, breast and belly gray brown, under tail coverts sooty brown; bare skin around eye, bill dark brown.

NESTLING: Undescribed. SOURCES: AMNH, BMNH, CM, FMNH, MVZ, ROM, UMMZ, USNM,YPM.

Subspecies Dasylophus superciliosus superciliosus (Dumont, 1823); larger, crest long, greatest length of red crest feathers 25–30 mm, breast darker;W and S Luzon; Dasylophus superciliosus cagayanensis (Rand and Rabor, 1967); smaller, crest short, greatest length of red crest feathers 10–12 mm, less white at base of crest feathers, breast paler; Cagayan Province, NE Luzon.

Measurements and weights D. s. superciliosus: Wing, M (n ⫽ 12) 143–160 (150.8 ⫾ 6.0), F (n ⫽ 9) 148–166 (156.5 ⫾ 7.0); tail, M 215–244 (229.0 ⫾ 8.1), F 213–235 (224.0 ⫾ 7.8); bill, M 33–39 (36.3 ⫾ 1.9), F 33–39.5 (36.7 ⫾ 2.0); tarsus, M 30–39.2 (33.3 ⫾ 2.7), F 32–38.3 (36.8 ⫾ 2.0) (AMNH,YPM); D. s. cagayanensis: wing, M (n ⫽ 4) 147–155 (150.5 ⫾ 4.1), F (n ⫽ 4) 140–151 (147.0 + 5.0); tail, M 210–224 (220.0 ⫾ 6.7), F 208–224 (213.5 ⫾ 7.2) (FMNH).

Scale-feathered Malkoha Dasylophus cumingi 309 Weight, D. s. superciliosus: M (n ⫽ 8) 100–129.9 (115.1), F (n ⫽ 12) 104–137 (119.6); D. s. cagayanensis, M (n ⫽ 2) 98.6–106 (102.3), F (n ⫽ 3) 95–102 (99.3) (FMNH, USNM). Wing formula, P6 ⬎ 5 ⬎ 4 ⬎ 7 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 38 cm. Long-tailed cuckoo with glossy black plumage and a bushy red eyebrow crest.

Voice Call, a soft guttural, slightly metallic sound “cheuk” given alone or repeated (Gonzales 1983, Scharringa 1999, Kennedy et al. 2000).

Range and status Philippine Islands: Luzon, Polillo, Marinduque and Catanduanes. Resident. On Mt Isarog, Luzon, the cuckoos occur between 300 and 760 m, and on Catanduanes they are common around Dugui-Too and in Mabini, Viga, along riverbanks. They are fairly common in forests, but these are being cut throughout the species range (Goodman and Gonzales 1990).

over bushes and grass, then dive headfirst to the ground and run (Gilliard 1950). Seen singly or in small groups, moving slowly through deep foliage and understory. Creep and hop in tangles of vines, and drop to ground to grab insects that fall from the branches (Delacour and Mayr 1946, Gonzales 1983).

Food Large insects (katydids, beetles), ants, worms, lizards and some plant material, probably also ground snails (Gonzales 1983, Goodman and Gonzales 1990).

Habitat and general habits

Breeding

Forests, both primary and second growth, wooded river-banks, creeks and lakes, and in grassland with bushes; from lowland to mountains, 100–800 m. They live near the ground and in the lower layer of the forest, somewhat like a coucal, in grass with bushes.When flushed they fly clumsily 10–20 m low

Season March to June, and a fledged chick was seen in Kangkoyawe, Caramoran, in June. Probably a nesting cuckoo with parental care, the nest and eggs are undescribed. The young fledge at 60 g (Gonzales 1983, Goodman and Gonzales 1990, Kennedy et al. 2000, FMNH).

Scale-feathered Malkoha Dasylophus cumingi (Fraser, 1839) Phoenicophaus Cumingi Fraser, 1839, Proceedings of the Zoological Society of London, 1839, p. 112. (Luzon, Philippine Islands) Other names: Lepidogrammus cumingi (Fraser, 1839) Monotypic.

Description ADULT: Sexes alike, head light gray, forehead whitish, crested crown and throat with flat scaly glossy blueblack feathers, unlike those of any other bird, the modified feather tip formed by barbs arranged linearly and the terminal barbs fused together and with

310 Scale-feathered Malkoha Dasylophus cumingi the rachis, upper back rufous brown, back black, wing, rump and tail black glossed dark green to blue, tail long with white tip; underparts, throat whitish, upper breast golden brown grading into rufous lower breast and darker on belly, lower belly and under tail coverts dull black, under wing coverts black; bare skin around eye red, iris red, bill yellow to pale creamy buff to brown, feet slate blue to gray.

pitched, a nasal “bläää”, a “raaab”, a light “prt” and a deep “krrr” which follows a higher tonal “krii-äääää” (Gonzales 1983, Raethel 1992, Scharringa 1999, Kennedy et al. 2000).

JUVENILE: Upper half and sides of head, neck, wing and body dark reddish brown, wing black, tail black with white tip as in adult, rectrices narrower than in adult; scaly head feathers grow while bird is still short-tailed; eye dark brown, bill dark brown.

Habitat and general habits

NESTLING: Undescribed at hatching, the feathered nestling is covered with dark rufous soft downy feathers and lacks the scaly ornamental head feathers of the grown juvenile, and the bill is dark. SOURCES: AMNH, BMNH, CM, CMNH, FMNH, MCZ, MVZ, ROM, UMMZ, USNM, YPM.

Measurements and weights Wing, M (n ⫽ 9) 150–169 (156.9 ⫾ 6.9), F (n ⫽ 12) 145–168 (159.0 ⫾ 6.3); tail, M 220–247 (241.4 ⫾ 15.2), F 220–245 (233.7 ⫾ 8.7); bill, M 30.5–41 (38.1 ⫾ 2.1), F 30.5–41 (36.5 ⫾ 3.0); tarsus, M 30.5–42 (34.9 ⫾ 3.1), F 34–39.5 (37.1 ⫾ 1.6) (AMNH, CMNH). Weight, M (n ⫽ 18) 148.5–204.3 (174.1 ⫾ 16.9), F (n ⫽ 24) 121.1–194.8 (167.2 ⫾ 19.8) (CMNH, FMNH, UMMZ, USNM,YPM). Wing formula, P6 ⬎ 5 ⬎ 4 ⬎ 7 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 42 cm. Long-tailed brown cuckoo with a whitish face, and a crest and throat with short, flat glossy blue-black feathers.

Voice An explosive high-pitched “quizzzzz-kid” or “whizzzzz-kid”, rising on the first note and leveling with the second note which ends with a snap, the call lasts about 1 sec. Calls are metallic and high

Range and status Philippines (Luzon, Marinduque, Catanduanes). Resident. Fairly common (Dickinson et al. 1991).

Forest, typically in primary and well developed secondary forest with vines and lianas, in replanted forest areas and in forest edge and second growth, from lowlands to 2000 m, on Mt Isarog between 300 and 900 m.They live in trees and vines, in low layers of the forest, from ground to as high as 12 m above the ground (Gilliard 1950, Gonzales 1983, Goodman and Gonzales 1990, Raethel 1992). Occur singly and in small groups, mostly skulking in dense understory foliage, or in the forest canopy where they take only short flights (Kennedy et al. 2000).

Food Insects, notably caterpillars (to 110 mm); centipedes, worms and small lizards and some plant material (Gonzales 1983, Goodman and Gonzales 1990). They takes birds out of a mist net, and in captivity they take insects and cooked meat (Raethel 1992).

Breeding March to May, estimated from dates of juveniles and ovaries of breeding females (Kennedy et al.

Chestnut-winged Cuckoo Clamator coromandus 311 2000, CMNH, FMNH).Although the nest is undescribed by biologists in the field, it is known by local people as a flat platform like the nest of a dove, built of twigs, grass and leaves, and located in a bush or more often high in a tree. Eggs are not described. The clutch size is 2 (Luzon, CMNH

RSK 3456, specimen label indicates 2 ovulated follicles; in captivity a female lays 2–3 eggs, Raethel 1992). Incubation and nestling periods are unknown in the field. In captivity a male built the nest and the pair incubated for 22 days, but the eggs did not hatch (Raethel 1992).

Genus Clamator Kaup, 1829 Clamator Kaup, 1829, Skizzirte EntwicklungsGeschichte und natürliches System der Europäischen Thierwelt . . . , 53. Old-World brood-parasitic cuckoos with a plumage crest. Type, by original

designation and monotypy, Cuculus glandarius (Linnaeus 1758). The name refers to the loud calling (L. clamare, to shout) of the bird. Four species.

Chestnut-winged Cuckoo Clamator coromandus (Linnaeus, 1766) Cuculus coromandus (Linnaeus, 1758), Systema Naturae, ed. 10, 171. (Coromandel) Other common names: Chestnut-winged Crested Cuckoo, Red-winged Crested Cuckoo Monotypic.

Description ADULT: Sexes alike, above glossy black, a long black crest, white collar, wing bright chestnut with no white patch, tips of primaries and secondaries broadly tipped dark brown, tail black with narrow white tips (T1 often worn with no remaining trace of white); underparts from chin to upper breast rufous, lower breast and upper belly white, lower belly gray, under tail coverts black; iris pale brown to dark brown, bill black with gray gonys, tarsi gray. JUVENILE: Above brown, barred with buffy edges of feathers, crown brown with buff tips, crest short, buff collar, buff edge to wing coverts, the bird appearing scaly rufous-buff above, tail tip and outer edge of feathers gray buff, underparts from throat to breast gray, belly whitish gray, under tail coverts dark gray.The crest appears at 5 weeks and the bird molts to adult plumage at 3 months (Osmaston 1916). NESTLING: Undescribed.

SOURCES: AMNH, BMNH, FMNH, MCZ, MVZ, MZB, ROM, TISTR, UMMZ, USNM, UWBM, ZRC.

Measurements and weights Wing, M (n ⫽ 11) 154–167 (161.3 ⫾ 4.0), F (n ⫽ 10) 155–165 (160.1 ⫾ 3.4); tail, M 216–247 (232.7 ⫾ 8.5), F 208–248 (230.6 ⫾ 14.5); bill, M 24–28.8 (25.8 ⫾ 1.5), F 22–26.6 (24.7 ⫾ 1.8); tarsus, M 22.5–26.6 (24.9 ⫾ 1.6), F 22–26.6 (23.8 ⫾ 1.71) (AMNH, UMMZ); Weight, M (n ⫽ 2) 65, (89.5, migrant, March), F (n ⫽ 2) 68, 77 (MZB, USNM); unsexed adults (n ⫽ 44) in Malay Peninsula were 66.4–83.8 (Wells 1999). Wing formula, P7 ⱖ 6 ⬎ 8 ⱖ 5 ⬎ 4 ⬎ 9 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 46 cm. A large crested cuckoo with white collar, rusty breast, chestnut wings and a long black tail. The juvenile is a subdued and mottled version of the adult.

Voice Loud, shrill, hoarse double whistle “klinck-klinck” is given repeatedly as a territorial call, in flight and when disturbed (B. King recording).The whistle is

312 Chestnut-winged Cuckoo Clamator coromandus similar to that of Mustached Hawk Cuckoo Cuculus vagans, with the paired notes less well spaced. Also harsh raucous screams or rapid grating “creech-creech-creech” or “critititit” (Smythies 1940, 1981, Ali and Ripley 1969, King and Dickinson 1975, MacKinnon and Phillipps 1993, Ali 1996, Grimmett et al. 1999, Robson 2000a).

Bishop 1997, Grimmett et al. 1999, Harrison 1999, Wells 1999, Robson 2000a, Sheldon et al. 2001). Vagrant in Korea, Japan, Okinawa, Taiwan and Hong Kong (King and Dickinson 1975,Wild Bird Society of Japan 1982, Park and Kim 1995, McWhirter et al. 1996, Thomas and Poole 2003). Uncommon.

Range and status

Habitat and general habits

Southern Asia from India (Himalayan foothills and terai zone from northern Uttar Pradesh east to Arunachal Pradesh and Assam in Khasi Hills and E Meghalaya), Nepal, Bhutan and NE Bangladesh to S and SE China including Hainan, and SE Asia in Burma, north and central Thailand, Laos, Tonkin, and north and central Annam. Seasonal in summer in northern India, China and northern Burma, where they arrive in May and depart in October. They are scarce residents in southern Burma (Tenasserim), Cambodia and southern Annam. They are passage migrants in SE India and the Malay Peninsula in autumn; they winter in India in the S Western Ghats, and winter in Sri Lanka, southeast Asia, the Malay Peninsula, the Malacca Straits to Sumatra, Lingga Archipelago, Mentawai Islands, Java, Borneo, Bangka, northern Sulawesi, Sulu Archipelago and the Philippines (Legge 1880, Oustalet 1899, Stresemann 1930b,Wildash 1968,Ali and Ripley 1969, van Merle and Voous 1988, Mees 1986, Smythies 1986, White and Bruce 1986, Cheng 1991, Dickinson et al. 1991, Coates and

Thick, low vegetation in secondary growth, broadleaved evergreen woodland, scrub and bushes, riverbank thickets, tall scrub bordering cultivation, bamboo thickets, oil palm, cultivated lands and villages; in lowlands and hills to 1500 m. In western Bengal, they occur mainly in the foothills and extend into the plains in forest areas (Inglis et al. 1920).They winter in similar habitats and in mangroves (Hoogerwerf 1948, Ali 1953, MacKinnon and Phillipps 1993, Coates and Bishop 1997, Grimmett et al. 1999, Wells 1999, Kennedy et al. 2000). Arboreal, they clamber in thick, low vegetation and in canopy foliage searching for insects, and they occur singly or in small loose groups.

Food Large insects, mainly caterpillars, beetles, mantids, ants; spiders (Sody 1989). A hand-reared fledgling developed its own feeding method: it severed the ends of a caterpillar, shook the caterpillar from

X

X X

X

Great Spotted Cuckoo Clamator glandarius 313 side to side to remove the gut, then it swallowed the caterpillar (Osmaston 1916).

Breeding In the rains. In India, they breed from March to August (Grimmett et al. 1999); in Sri Lanka from June to October (Legge 1880). Brood-parasitic. Hosts are mainly laughingthrushes Garrulax spp., often Greater Necklaced Laughingthrush G. pectoralis and Lesser Necklaced Laughingthrush G.

monileger (Osmaston 1916, Mackenzie 1918, Baker 1942, Becking 1981). Eggs are pale blue, unmarked, with a dull surface, 27 ⫻ 23 mm, larger than the egg of Indian Jacobin Cuckoo C. jacobinus (Baker 1906, 1934, 1942, Becking 1981); 1 to 4 cuckoo eggs are laid in a nest. Incubation and nestling periods are unknown. Cuckoo nestlings do not evict the host eggs, but they crowd out the host nestlings, which are trampled and starve in the nest (Osmaston 1916).

Great Spotted Cuckoo Clamator glandarius (Linnaeus, 1758) Cuculus glandarius Linnaeus, 1758, Systema Naturae, ed. 10, 111. (Northern Africa and southern Europe ⫽ Gibraltar) Monotypic.

Description ADULT: Sexes alike, crown gray with fine black shaft streaks and a short crest, back brown with white edges or tips on the feathers, wing coverts brown with white spots, wing flight feathers gray brown, tail dark gray-brown, inner feathers T1 dark throughout the length except for a small white tip in fresh plumage, tips of outer feathers T2 to T5 with large white tips, cheek gray with a darker streak extending from under the eye to the nape; underparts, throat buff, breast to belly and under tail coverts white; eye-ring gray to reddish, iris brown, bill gray to black, feet gray to black. JUVENILE: Crown and face black, back brown with white spots smaller than in the adult, wing coverts brown with small white spots, flight feathers rufous, tail rufous, white spots on tips of tail feathers smaller (length ⬍ 10 mm, vs ⬎ 20 mm in adult), underparts with chin and throat a darker buff than in adults; eye-ring orange while bird is growing its feathers, iris dark brown, bill dark gray, feet dark gray. First-year birds often have a mixed-age plumage of old rufous and new gray-brown wing feathers, with an incomplete transilient molt in which they replace alternate primaries, so the wing has an alternating sequence of brown and rufous primaries.

NESTLING: Naked, yellow or pink at hatching.The mouth lining is pink with whitish spicules on the palate (Valverde 1971, UMMZ photographs). SOURCES: AMNH, BMNH, BWYO, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM.

Subspecies and history Clancey (1951) described Clamator glandarius choragium as a smaller subspecies in southern Africa. Although birds breeding in southern Africa are smaller on average, they overlap in size with birds in the Mediterranean and it is not possible to distinguish between northern and southern birds where both might occur, as in Kenya during the northern nonbreeding season.

Measurements and weights Mediterranean: Wing, M (n ⫽ 11) 194–215 (210.2 ⫾ 6.5), F (n ⫽ 7) 191–218 (202.7 ⫾ 9.9); tail, M 192–230 (218.7 ⫾ 11.1), F 192–211 (203.4 ⫾ 5.9); bill, M 23–28 (25.1 ⫾ 1.4), F 22–28 (25.3 ⫾ 2.9); tarsus, M 28–34 (31.0 ⫾ 2.0), F 30–34 (31.5 ⫾ 1.5) (AMNH, MSNG, MSNM); South Africa:Wing, M (n ⫽ 11) 184–210 (199.0 ⫾ 7.0), F (n ⫽ 9) 186–202 (194.1 ⫾ 8.0); tail, M 204–230 (211.8 ⫾ 9.0), F 183–210 (194.9 ⫾ 7.8); bill (head) M 22–24 (23.0 ⫾ 0.7), F 22, 22; (nostril), M 17–21 (19.6 ⫾ 1.4), F 17–19 (17.8 ⫾ 0.8); tarsus, M 28–32 (30.0 ⫾ 1.7), F 27–33 (29.0 ⫾ 1.9) (MVZ, UMMZ, USNM). Weight, South Africa, M (n ⫽ 10) 128–168 (139.3), F (n ⫽ 9) 118–148 (128.7) (UMMZ).

314 Great Spotted Cuckoo Clamator glandarius Wing formula, P8 ⫽ 7 ⱖ 6 ⬎ 9 ⱖ 5 ⬎ 4 ⱖ 3 ⬎ 2 ⬎ 1 ⱖ 10.

Field characters Overall length 39 cm. A large, noisy short-crested cuckoo with long wings and long graduated tail, gray above with bold white spots on the wing, and white below with a buff breast.The adult has a gray crown and gray brown wings, the juvenile has a black crown and cheeks and rufous wings, and the first-year bird has a mix of these colors.

Voice Harsh guttural “gah-gah-gah . . . gak-gak-gak . . . ko-ko-ko-ko”, the series falling in pitch and increasing in tempo, a trilled “kirrrrr” and cackled notes (Chappuis 1974, Cramp 1985). Male gives call both perched and in flight. In courtship when the pair is together, a continuous rapid and soft chittering. Female calls when she locates a host nest, a repetitive “woing”.

Range and status Southern Europe from Iberian Peninsula, Balearic Islands, southern France, Italy (not known as breeder), Croatia (local on coast), Greece, Bulgaria, Turkey, Iraq, the eastern Mediterranean, Morocco,

Cyprus and Egypt (Cramp 1985, Maumary and Duperrex 1991, Boet and Boet 1996, Dias et al. 1996, Shirihai 1996, Hagemeijer and Blair 1997). In sub-Saharan Africa they breed from Senegal, Ghana, Mali, Nigeria, Sudan, Eritrea, Ethiopia, northern Somalia and Kenya, south to Tanzania, Angola and South Africa (Irwin 1988). In Spain, adults occur mainly from April to June and leave the breeding area in mid June, juveniles leave between early July and early August with some staying until November (Hagemeijer and Blair 1997); in Guadalquivir River area they increase in numbers from December to March, suggesting an early return in winter (Herrera and Hidalgo 1974). In Israel the breeding birds arrive in late winter and adults leave by June, the juveniles in June and July (Shirihai 1996). European birds migrate to Africa mainly north of 10° N in West Africa. Hundreds appear in late summer in southern Mauritania and later in Senegal (Moreau 1972). Other birds in Senegal are resident and the two populations do not differ morphologically (Morel and Morel 1990). In The Gambia juveniles appear irregularly from the north, and adults occasionally breed there (Barlow et al. 1997). In West Africa they breed as migrants (Giraudoux et al. 1988, Elgood et al. 1994). In Somalia they occur mainly from

Great Spotted Cuckoo Clamator glandarius 315 November to June (Ash and Miskell 1998), in Eritrea they appear in coastal acacias from December to March (Moreau 1972). In Arabia, occasional migrants occur in season and the species is rare in winter ( Jennings 1981, Goodman and Jennings 1988, Porter et al. 1996). In East Africa birds occur all year, and migrants occur in passage across the Serengeti plains from January to March (Schmidl 1982, Zimmerman et al. 1996, Stevenson and Fanshawe 2002). In Kenya the birds occur mainly from October to March; some, perhaps from breeding grounds in southern Africa, are very fat with small ovaries in May, when other fat birds are laying (MVZ). Local movements occur in relation to rains and aridity. In southern Africa the cuckoos are mostly migratory. In Zambia they occur from October to April (Aspinwall and Beel 1998), and as far south as the thornveld in Natal and Cradock in the Eastern Cape of South Africa a few remain all year, and birds live in groups as large as six in July (Payne 1969a, Cyrus and Robson 1980); they are more common in the austral breeding season from October to December.Vagrant in Britain, northern Europe, and Russia. In Britain and Ireland, first-year birds appear from February to October (Lévêque 1968, Lansdown 1995). Generally uncommon, they are conspicuous; in southern Africa they are most numerous in the acacia thornbush of the Eastern Cape. Most of the Palearctic population is in Spain, with 98% of the European population.The species has expanded its range in Spain and the Middle East. It has decreased in numbers in Egypt where it is now rare as a breeding bird; it formerly bred in Algeria, and there are a few records suggesting breeding in southern Europe from Italy to the Balkans (Lévêque 1968, Cramp 1985, Isenmann and Moali 2000). Density in optimal habitat in southern Spain, 3–9 birds/km2. In Spain the estimated population is 55,000–64,000 pairs; in Turkey it is 1000–5000 pairs (Dias et al. 1996, Hagemeijer and Blair 1997).

Habitat and general habits Semi-arid open woodland especially thorn-scrub acacia, pastures, other scrub, almond groves, olive groves, thickets, rocky hillsides in dry savannas;

mainly between 500–2000 m. In Spain, they live in wooded areas with pines Pinus spp. and oaks Quercus ilex (Dias et al. 1996, Hagemeijer and Blair 1997).They feed in trees taking caterpillars, perching and searching by scanning the foliage. The cuckoos often feed on the ground, taking caterpillars and grasshoppers, and sometimes remove hairs from the caterpillars before eating them, at other times leaving the hairs on their food; the hairs then line their stomach in a felt (Winterbottom 1962, RBP).

Food Insects, mainly large hairy caterpillars and grasshoppers, crickets, beetles, moths and termite alates, dragonflies; small lizards (Di Carlo 1971, Rowan 1983, Cramp 1985). In Spain its numbers vary with the abundance of hairy caterpillars of the processionary moth Thaumetopoea pityocampa (Hagemeijer and Blair 1997).

Displays and breeding behavior Territorial in breeding season, the area ranges from 1 to 8–9 km2; in Spain, averages c. 3.7 km2, an area with c. 60 host nests. Male advertises territory with flight and calls, rising to about 20 m with rapid wingbeats, descends while calling with its tail fanned, and glides along just above the ground.The large territories are not completely exclusive. More than one female sometimes lays in the same area and in the same nest, as shown by genetic differences between the eggs. On the other hand, a female sometimes lays more than one egg in a nest (Arias de Reyna 1998, Martínez et al. 1998b). A female lays on alternate days, in the first hours of the morning when the host is away from the nest. She lays eggs in sets, as many as 6 eggs over 12 days, or 12–18 eggs in a season. Females breed in firstyear plumage when they have a few retained juvenile flight feathers and breed as older adults in full adult plumage (Mundy and Cook 1971, Payne 1973a, 1974, Arias de Reyna 1998). In courtship, the male offers the female a large insect and the pair mates. Courtship feeding occurs at other times as well, and occasionally an adult cuckoo feeds a fledged young. In mating, the male mounts the back of the crouched female and passes

316 Great Spotted Cuckoo Clamator glandarius food to her bill, then each bird holds an end of the insect, and the male releases the food when copulation is complete. The pair copulates several times in an hour (von Frisch 1969, Pettet 1975, Soler 1980, Cramp 1985). Male and female cuckoos cooperate in laying when the male distracts the host from the nest and the female enters the nest and lays (von Frisch 1969, 1973, Alvarez and Arias de Reyna 1974).The female sometimes visits the nests alone, perhaps inspecting the laying status of the hosts. A female that peered into nest holes in a colony of Pied Starlings Spreo bicolor in the Eastern Cape was in breeding condition but was not laying and had not laid for a few days (UMMZ).

Breeding In the Mediterranean region from late April to early June (Cramp 1985, Arias de Reyna 1998). In Egypt they breed in March (Cramp 1985, Goodman and Meininger 1989) and in southern Morocco in March (Maumary and Duperrex 1991). In sub-saharan Africa they breed mainly during early rains, in Senegal in April and May, and from The Gambia to Nigeria they breed from April to July (Bannerman 1951, Grimes 1987, Giraudoux et al. 1988, Gore 1990, Morel and Morel 1990, Elgood et al. 1994). In NW Somalia they breed from late April to June (Ash and Miskell 1998). In East Africa the time of breeding varies with local rainy seasons, east of the rift valley in April and May and October to December and in the arid northwest from February to April (Brown and Britton 1980, MVZ). In southern Africa they breed in Malawi in October and November (Benson and Benson 1977), in Angola in October (Dean et al. 1988), in Namibia and Botswana from December to March, and in South Africa from October to January; locally the breeding season is about 10 weeks (Payne and Payne 1967, Payne 1973a, Rowan 1983, Irwin 1988, Maclean 1993, Skinner 1996). Brood-parasitic. Hosts are crows and magpies in the Mediterranean region, particularly Magpie Pica pica, Azure-winged Magpie Cyanopica cyanea and Carrion Crow Corvus corone, and occasionally Jackdaw C. monedula and Red-billed Chough

Pyrrhocorax pyrrhocorax in Spain (Zuñiga et al. 1983, Cramp 1985, Soler 1990), and Hooded Crow C. corone in Israel (Shirihai 1996) and Egypt (Goodman and Meininger 1979). In Israel they also parasitize House Crow Corvus splendens which was first reported there in 1976 (Shirihai 1996), probably from its closest established population in Egypt on the Red Sea (Goodman and Meininger 1989). The cuckoos have successfully parasitized the crow at least twice (1997, 2002), and this is the first time these two species are known to have been in contact (Yosef 1997, 2002). In Africa the Great Spotted Cuckoos parasitize Pied Crow C. albus, Cape Rook C. capensis, Fantailed Raven C. rhipidurus and Dwarf Raven C. edithae (Friedmann 1964, Irwin 1988). In West Africa they may also parasitize hole-nesting starling, as fledgling cuckoos in Ghana were mobbed by Purple Glossy Starling Lamprotornis purpureus (Grimes 1987). In East and southern Africa they parasitize both crow nests in trees and starling nests in holes in trees (glossy starlings Lamprotornis spp., notably Burchell’s Glossy Starling L. australis, Greater Blue-eared Glossy Starling L. chalybaeus and Cape Glossy Starling L. nitens), rocky cliffs (red-winged starlings Onychognathus spp.) and tunnels in eroded sandbanks (Pied Starling Spreo bicolor) (11 species of starlings are known as hosts, Feare and Craig 1999) (Friedmann 1948, 1956, 1964a, Payne and Payne 1967, Jensen and Jensen 1969, Trevor and Lack 1976, Brown and Britton 1980, Schmidl 1982, Nikolaus 1987, Dean et al. 1988, Irwin 1988,Ash and Miskell 1998).The most commonly parasitized hosts in South Africa in southern Natal and the Eastern Cape are Pied Starling and Red-winged Starling Onychognathus morio, and in Namibia in the rocky escarpment region west of Windhoek the main host is Pale-winged Starling O. nabouroup (Kemp et al. 1972, Rowan 1983). In acacia thornveld in Weenen County, Natal, South Africa, the cuckoos parasitize two kinds of crows and three kinds of native starlings as well as introduced Indian Mynah Acridotheres tristis (Symons 1962,West et al. 1964, Sharp 1976). Eggs are pale greenish with brown and gray marks, 31 ⫻ 23 mm (Soler 1990), smaller than crow eggs and slightly larger than starling eggs. Eggs in

Great Spotted Cuckoo Clamator glandarius 317 Spain do not vary with the size or color of host eggs; in both Carrion Crow nests and Magpie nests the cuckoo eggs resemble Magpie eggs. In subSaharan Africa the eggs may be less densely spotted than eggs in the Mediterranean (Cramp 1985, Irwin 1988). The cuckoo eggs in nests of Pied Starling, the most common host in parts of eastern South Africa, are spotted, not uniformly colored like the starling eggs (Plate 17). Often several cuckoo eggs are laid in a nest. In the Mediterranean most parasitized crow nests have a single cuckoo egg (32 of 42 nests), and in southern Africa about half of the parasitized crow nests have 2 or more cuckoo eggs, range 2–4 (Friedmann 1964a); a nest in Natal had 9 cuckoo eggs (West et al. 1964). Nests with two or more cuckoo eggs are sometimes parasitized by one female, sometimes by two females (Martínez et al. 1998b). Females lay while perched on the nest edge, dropping an egg from the cloaca into the clutch, and most parasitized nests have host eggs that are damaged, crushed or cracked by impact (eggs in unparasitized nests are not damaged in this way) (Soler 1990). Incubation period averages 12.8 days, shorter than in the host. Nestling cuckoos do not evict. They hatch at 8 g, reach 50% of fledging weight 8 days after hatching, and grow steadily to day 14 when growth levels off. Begging young raise the neck upward with the mouth open, unlike Common Cuckoo Cuculus canorus; they crouch when begging intensity is low, and on rare occasions they flutter their wings. Sometimes the young produce foul feces, in contrast to the host nestlings and to other brood-parasitic cuckoos (Cuculus, Chrysococcyx). They fledge at 110 g in 22–28 days (before the host young fledge), then the foster parents feed them for another 25–59 days (von Frisch and von Frisch 1967, Valverde 1971, Alvarez and Arias de Reyna 1974, Soler and Soler 1991, Arias de Reyna 1998, M. Soler, in litt.). In Magpie nests with more than one cuckoo, the younger cuckoo

may starve; the first and second cuckoo eggs are more successful than later eggs in hatching and fledging. In crow nests more than one cuckoo chick often fledges. In some nests of starlings, holenesting birds that are often multiply parasitized with cuckoo eggs, more than one cuckoo survives (Red-winged Starlings: Feare and Craig 1999).The begging call of nestling cuckoos is like that of certain hosts, with different calls of cuckoos in nests of crows and Magpies, and cuckoos in nests of crows and starlings (Redondo 1993). Cuckoo parasitism affects host nesting success when the female removes host eggs from the nest (when the cuckoo lays after the hosts have incubated), when she damages the host eggs as she lays and her egg drops onto the clutch and cracks the host eggs, when she pecks at host eggs, and when the cuckoo nestlings compete with host nestlings for food and parental care (Valverde 1971, Mundy and Cook 1974, 1977, Soler 1990, Soler and Soler 1991, Soler et al. 1996, 1997). Young cuckoos receive parental care for nearly five weeks after fledging, compared to three weeks for young Magpies (Soler et al. 1994a). Magpie chicks grow slowly and fledge at low weight in parasitized nests due to cuckoo chicks out-competing or smothering them, and few or none of these host young fledge; the Magpie chicks die when they hatch 3–4 days later than the cuckoo. In contrast, cuckoo nestlings do not affect the success of young crows, which are larger. Success of cuckoos from laying to fledging varies with the time when a cuckoo egg is laid in the nest, with 92% success when the cuckoo egg is laid when the host has only 1–3 eggs, to 10% success when the host clutch is complete. Survival of cuckoo young to independence is 63%, (Arias de Reyna et al. 1982). Great Spotted Cuckoos have bred in captivity, parasitizing a nesting pair of Hoopoe Upupa epops (Mendelsson and Golani 1986).

318 Levaillant’s Cuckoo Clamator levaillantii

Levaillant’s Cuckoo Clamator levaillantii (Swainson, 1829) Coccyzus levaillantii Swainson, 1829, Zoological Illustrations, 2nd series, 1, pl. 13 (Senegal) Other common names: African Striped Cuckoo, Striped Cuckoo, Striped Crested Cuckoo. Other names: Oxylophus levaillantii (Swainson), Clamator cafer (Lichtenstein). Monotypic.

(24.1 ⫾ 1.1), F 23.1–26.4 (24.3 ⫾ 1.2); tarsus, M 28–32 (30.1 ⫾ 1.5), F 28–32.3 (30.6 ⫾ 1.4) (AMNH, UMMZ). Weight, M (n ⫽ 5) 111–145 (128.6), F (n ⫽ 3) 115.7–125 (120.9) (AMNH, UMMZ). Wing formula, P7 ⱖ 6 ⬎ 8 ⱖ 5 ⬎ 9 ⫽ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Description

Field characters

ADULT: Sexes alike, above glossy black with long black crest, wing black with white patch, tail black with white tips; underparts white broadly streaked with black on throat and breast, and finely streaked with black on lower breast, flanks and leggings; iris brown, bill black, feet blue-gray. In coastal East Africa a dark phase occurs with the underparts black and the wing patch and tail tip white.

Overall length 39 cm. Large crested cuckoo, black above and white below, broadly streaked black on throat and breast, and streaked leggings (unstreaked in C. jacobinus). Juveniles in most of the range are brown and white, with a short crest and white spots in the wing.

JUVENILE: Above dark brown to rufous, wing black with white patch, tail gray with paler tip; underparts whitish with dull brown on breast, the breast sometimes with black streaks; iris pale gray to dark brown, gape bright red-orange, papillae on either side of palate and at base of tongue paler orange. A few juveniles in the eastern Congo (Beni (Bungulu), Ituri, AMNH 262583; Poko, Uele, BMNH 1919.10.12.152) have uniformly rufous upperparts and underparts including the breast and under tail coverts, and the dark brown wing lacks the white patch.

Call a ringing “kuwu-weer, kuwu-weer . . .”, often ending with rapid “kwi’kwi’kwi’kwi . . .” similar to Jacobin Cuckoo C. jacobinus’, but longer, louder and harsher (Chappuis 1974, 2000) and the “weer” notes with a burr on the end. They also give a squealed single “kuweer”. The begging call of the young differs from that of the host’s young in some observations (Vernon 1982); in others, the fledged cuckoo calls “ker, ker, ker” like a begging young babbler (Vincent 1946, Jubb 1952, Mundy 1973, Steyn 1973, Stjernstedt 1993, Hustler 1997b, Barry 1998).

NESTLING: Naked at hatching, skin pink to pinkorange; eye-lids yellow, upper mandible black with protuberant oval nostrils, gape red-orange with no fleshy swellings at the side.

Range and status

SOURCES: AMNH, BMNH, BWYO, CM, DMNH, FMNH, MSNG, ROM, UMMZ, USNM.

Measurements and weights Wing, M (n ⫽ 11) 167–184 (173.9 ⫾ 5.8), F (n ⫽ 10) 166–180 (174.2 ⫾ 4.7); tail, M 178–248 (217.1 ⫾ 18.2), F 192–233 (221.1 ⫾ 11.7); bill, M 22.8–25.9

Voice

Africa south of the Sahara, breeding from Senegal to East Africa (Eritrea, Ethiopia and Kenya), central and southern Africa (Angola, N Namibia and Zambia to Transvaal and N Natal); nonbreeding birds occur to central and coastal West Africa. IntraAfrican migrant; seasonal in southern Africa, resident or partly migratory in equatorial regions. In West Africa they breed in northern woodlands in the rains, then move closer to the coast into southern woodlands and the forest zone where they molt in the dry season (Bannerman 1951, Brunel 1958, Lamarche 1980, Louette 1981, Grimes 1987, Morel

Levaillant’s Cuckoo Clamator levaillantii 319 mopane Colophospermum mopane woodlands, in more moist habitat than Jacobin Cuckoo.

Food Insects, mainly hairy and spiny caterpillars, including mopane worms Imbasia belina (Styles 1995), winged termites, beetle larvae and grasshoppers; occasionally take berries and young shoots from trees (Brooke 1985b).

Displays and breeding behavior

and Morel 1990, Elgood et al. 1994, Barlow et al. 1997, Gatter 1997, Anciaux 2000). They are common and conspicuous in the rains in The Gambia and Mali but disappear in the dry season, and are rainy season visitors in Sudan (Bannerman 1951, Nikolaus 1987). In Gabon and (one record) Bioko, nonbreeding birds appear in the dry season in December (Christy and Clarke 1994, Pérez de Val et al. 1997). In East Africa they come and go with the rains. In southeast Tanzania to Uganda they occur from October to June, in western Kenya from May to September, and the black-phase birds are present mainly in northeastern Tanzania and coastal Kenya from May to November (Stevenson and Fanshawe 2002) or March to September (Zimmerman et al. 1996) and were once seen in Somalia in April (Ash and Miskell 1998). In southern Zaire (Upemba NP) and Zambia, they occur from October to April (Verheyen 1953, Aspinwall and Beel 1998). In southern Africa they arrive in October-November and depart in March-April or later in Zimbabwe (Rowan 1983); southern birds are suspected to migrate to equatorial and West Africa. Population density, c. 1 calling bird /150 ha in Transvaal (Tarboton et al. 1987).

Habitat and general habits Open woodland, thickets, gardens, riverine forest, scrub (Rowan 1983), in E Africa mainly in areas with more than 500 mm rainfall (Lewis and Pomeroy 1989), in Zimbabwe also in Acacia-Terminalia and

Often seen in breeding pairs that move and feed together. Two birds face each other, calling, with crests raised and heads bobbing, wings half open and tails fanned; it is uncertain whether this is courtship or aggressive behavior (Lorber 1984). In courtship, one adult mounts the other and feeds it a caterpillar (Leonard 1998b). The pair works together in parasitizing a nest, the male drawing the host away from the nest, while the female cuckoo enters the nest to lay (Cackett 1970). A female lays about 22 eggs in a breeding season (Payne 1973a, 1974).

Breeding During the rains, in The Gambia from June to November (Gore 1990), in Ghana and Togo mainly February to June (Grimes 1987, Cheke and Walsh 1996), in Nigeria from April to September (Serle 1977, Elgood et al. 1994), in Ethiopia from June to September (Urban and Brown 1971), in Kenya they breed from March to May (Brown and Britton 1980), in Angola from September to March (Dean 2000), in eastern Zaire there are nestlings in May and October (Chapin 1939), in Malawi and Zimbabwe they breed from December to April with a few young appearing as late as June (Benson and Benson 1977, Irwin 1981, 1988,Vernon 1982), and in Botswana they breed from January to May (Skinner 1996). Brood-parasitic. Hosts are mainly babblers Turdoides (Brown Babbler T. plebejus, Black-capped Babbler T. reinwardii, Arrow-marked Babbler T. jardineii,White-rumped Babbler T. leucopygia, Barecheeked Babbler T. gymnogenys, Southern Piedbabbler (Pied Babbler) T. bicolor, Capuchin Babbler Phyllanthus atripennis), and fledged young are fed by

320 Jacobin Cuckoo Clamator jacobinus Chestnut-bellied Starling Spreo pulcher (Chapin 1939, Friedmann 1948, 1956, 1964a, West et al. 1964, Lamarche 1980, Moore 1984, Jones 1985, Nikolaus 1987, Gore 1990, Jones 1992, Hustler 1997b, Barry 1998, Leonard 1998, Dean 2000; Allan 2002). Eggs are unmarked greenish blue (Chapin 1939, Vernon 1982, Irwin 1988, Tarboton 2001); in Nigeria cuckoo eggs in nests of babblers, pink with darker speckles, like local babbler eggs (Serle 1977, Sharland and Serle 1977, BMNH). Egg color matches the host eggs, though the surface texture is more pitted than in babbler eggs (Vernon 1982). The female cuckoo lays up to 4 eggs in a series; she ovulates and lays on alternate days and carries the egg in the oviduct for a day (Payne 1973a). Egg size, 26 ⫻ 21 mm (Vernon 1982), pink eggs in Nigeria (n ⫽ 2) 24.5 ⫻ 20.4, 25.7 ⫻ 20.8 mm (BMNH). The embryo develops for a day in the oviduct (Steyn and Howells 1975). The incubation period is 11 days (Steyn and Howells 1975) to 12–13 days ( Jones 1985, 1992)

and the nestling period in one case was 10 days (Jones 1985). The nestling skin turns blackish by day 5 and spike-like quills appear by day 7. The young grow rapidly, from 5–6 g at hatching to 52 g on day 6 (Steyn and Howells 1975), or to 34 g on day 5 and 49 g on day 7 (Steyn 1973).The nestling cuckoo does not evict the host eggs or young, and more than one young babbler sometimes fledges from a nest where a cuckoo fledges (Vernon 1982, Rowan 1983, Jones 1985, 1992, Hustler 1997b). Adult cuckoos sometimes visit the nest when a cuckoo nestling is fed by the foster parents ( Jones 1985). After fledging, the young cuckoo begs with wing flutter and gives a begging call like that of the host young; it is fed by foster parents for at least 36 days. At times the fledgling is mobbed by the host group, then when it begs it is fed by the host group (Vincent 1934, Mundy 1973,Vernon 1982, Rowan 1983, Barry 1998). In southern Africa about 7.5% of Arrow-marked Babbler nests are parasitized by the cuckoo (Payne and Payne 1967).

Jacobin Cuckoo Clamator jacobinus (Boddaert, 1783) Cuculus jacobinus (Boddaert, 1783), Table des Planches enluminéez d’histoire naturelle, p. 53 (Coromandel Coast) Polytypic. Three subspecies. Clamator jacobinus jacobinus (Boddaert, 1783); Clamator jacobinus pica (Hemprich and Ehrenberg, 1833); Clamator jacobinus serratus (Sparrman, 1786). Other common names: Pied Cuckoo, Pied Crested Cuckoo, Black Crested Cuckoo, Black and White Cuckoo. Other names: Oxylophus jacobinus (Boddaert), Clamator serratus (Sparrman).

JUVENILE: Above dull brown to sooty black, crest short (⬍ 15 mm, feathers rounded, vs pointed in adult), pale-edged wing coverts, the wing patch sometimes off-white or light brown and smaller than in adult, tail spots buff not white and with the spots V-shaped not truncate, outer vane of outer tail feather T5 buffy white, tail feathers narrower than in adult (15–25 mm, vs 28–36 mm); underparts whitish with throat gray and breast buffy, a pale buff-white malar stripe between black cheek and gray throat; iris pale yellow, bill with lower mandible yellow at base. A rare all-rufous color phase occurs in Central Africa.

Description ADULT: Sexes alike, above glossy black with a black crest, wing black with a large white patch, tail long and black, central tail feather T1 black with narrow white tip, outer tail feathers T2 to T5 with large white truncate tips; underparts, throat to under tail coverts white (underparts black in a color phase in South Africa); iris brown, bill black, feet slate gray.

NESTLING: Naked at hatching, skin pink to orange pink, darkening to purplish brown by 48 hours; mouth lining red, sides of gape yellow. SOURCES: AMNH, BMNH, BWYO, CM, FMNH, MCZ, MSNG, MVZ, ROM, UMMZ, USNM.

Jacobin Cuckoo Clamator jacobinus 321

History and subspecies Clamator jacobinus jacobinus (Boddaert, 1783); smaller, throat and upper breast white; south of 15°N in southern India and Sri Lanka; Clamator jacobinus pica (Hemprich and Ehrenberg, 1833); larger, throat and upper breast white or white with hair streaks; southern Asia to Pakistan, northern India, Nepal, S Tibet, Kashmir and Burma, and Africa south to Tanzania and Zambia; Clamator jacobinus serratus (Sparrman, 1786); size as in C. j. pica, two color phases in both males and females, one phase below white-washed gray with thin blackish streaks on throat; the other phase all black with a white wing patch and no white tail patch in coastal South Africa; Zambezi River to the Cape region of South Africa. Although two subspecies are recognized in India, the birds in northern India are not consistently larger than those in southern India and Sri Lanka. Stresemann (1924c) first recognized the two color phases in southern Africa as the same species.

Measurements and weights C. j. pica, India north of 20°N: Wing M (n ⫽ 17) 143–154 (147.2 ⫾ 2.8), F (n ⫽ 26) 140–156 (148.5 ⫾ 3.9); tail, M 160–174 (167.4 ⫾ 6.1), F 155–173 (161.2 ⫾ 5.7); bill, M 19–23 (21.3 ⫾ 1.4), F 19–22 (20.08⫾1.1); tarsus, M 23–25 (24.0 ⫾ 0.8), F 22–26 (23.7 ⫾ 1.1) (AMNH, BMNH, CU, FMNH, USNM, UMMZ); Arabian Peninsula (April–June): Wing, M (n ⫽ 5) 148–158 (154.5), F (n ⫽ 5) 144–168 (157.8) (AMNH, BMNH); C. j. jacobinus, India south of 15°N and Sri Lanka: Wing, M (n ⫽ 5) 140–148 (145.5 ⫾ 2.2), F (n ⫽ 6) 137–160 (143.5 ⫾ 8.8) (AMNH, BMNH, FMNH); C. j. serratus, South Africa: Wing, M (n ⫽ 12) 146–167 (153.9 ⫾ 5.4), F (n ⫽ 12) 148–157 (152.4 ⫾ 3.1); tail, M 178–192 (183.8 ⫾ 3.9), F 173–186 (179.0 ⫾ 4.4); bill, M 19–22 (20.8 ⫾ 0.8), F 20–22 (10.6 ⫾ 0.7); tarsus, M 24–27 (25.3 ⫾ 1.2), F 24–28 (25.9 ⫾ 1.1) (UMMZ). Weight, C. j. serratus, South Africa: M (n ⫽ 11) 69.2–87.6 (78.6 ⫾ 5.0), F (n ⫽ 14) 80.9–104.0 (90.1 ⫾ 6.5) (the heavier females had an egg in the oviduct) (UMMZ).

Wing formula, P7 ⬎ 8 ⬎ 6 ⬎ 9 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 10 ⬎ 1. Comments: although the birds of southern India are small, so are some birds as far north as the Ganges Plain. On average the southern birds are smaller, and in this case the behavior in migration and the wintering regions of the northern and southern birds supports the idea that the two populations differ.

Field characters Overall length 34 cm. Long-tailed crested cuckoo, black above and white below, the wing and tail with a white patch, birds unstreaked in India and northern Africa, legging unstreaked, underparts sparsely streaked black (or all black with white wing patch) in southern Africa.

Voice Loud ringing “kleeuw, kleeuw, kleeuw”, each note descending, repeated in series, given by both sexes. Males often follow the series with a faster series of short rising notes “kwik-kwik-kwik . . .!”. Call, a loud “kweek!” and a squealed “kwir’kik”. Calls are the same in southern Africa, East Africa and India (Chappuis 1974, 2000, White 1984, Irwin 1988, Stjernstedt 1993, Ali 1996).

Range and status In Africa, the cuckoos are resident in tropical areas and intra-African migrants at higher latitudes above 18°N and S. In Asia they breed in southeast Iran, the Indian subcontinent and Burma. In both regions they arrive on the breeding grounds with the seasonal rains. In West Africa they breed in Mauritania and migrate to Senegal and The Gambia in November and December (Gore 1990, Morel and Morel 1990, Barlow et al. 1997). In Mali they occur from May to October; in Ivory Coast from February to April when not breeding (Thiollay 1985); in Togo and Nigeria seasonal in the north in April– September and in the south in the dry season (Elgood et al. 1994, Cheke and Walsh 1996); in savanna and forest edge in Gabon, juveniles appear as rare migrants in December (Christy and Clarke

322 Jacobin Cuckoo Clamator jacobinus

1994). In Eritrea and Somalia they appear from April to August and October to December, including a mix of wintering birds from southern Africa, wintering birds from India and breeding birds of East Africa (Ash and Miskell 1998, MSNG). In East Africa C. j. pica seen to move south from November to January (Stevenson and Fanshawe 2002); in the Serengeti they move northwest in migration over the plains in February and March (Zimmerman et al. 1996), while in Uganda they move north at Nimule in March and April and south along Lake Albert in November ( Jackson 1938). In southern Zaire they appear from the end of September to March, in Zambia mainly from November to March, in the Eastern Cape they appear in late October and disappear after December when they move to tropical regions for the dry season (Verheyen 1953, Skead 1995, Vernon et al. 1997, Aspinwall and Beel 1998). Black-phase C. j. serratus breed in the southern Cape Province (Eastern Cape to southern W Cape) and Natal.They appear in migration from Transvaal to Zambia and Malawi, and in austral winter from April to September north to Zanzibar, Kenya, Ethiopia, Sudan, Chad, Gabon and Nigeria (Friedmann 1930, 1964a, Bannerman 1951, Clancey 1960, Irwin and Benson 1966, Benson

et al. 1970, Benson and Benson 1977, Pakenham 1979, Irwin 1988, Stevenson and Fanshawe 2002); they are quiet on their wintering grounds. The black-phase bird in Gabon was in fresh plumage in November, perhaps a late southward migrant. Jacobin Cuckoos are seasonal in Pakistan where they arrive in June before the monsoon rains and they depart in autumn, and occur in the foothills of the Himalayas, near Islamabad, and along wooded areas of the Indus plains (Biddulph 1881, Roberts 1991). In northern India, C. j. pica occur from April or May to September and November and breed in the plains (Inglis et al. 1920, Whistler 1928, 1931). In the breeding season, they are common north to Bihar and the terai and duars, and even seen rarely in the Himalayas as high as 4200 m (Ali and Ripley 1969, Jamdar 1987). In Nepal and extreme southern Tibet they are uncommon, in Pakistan common, in Bangladesh local (Grimmett et al. 1999); in Burma they are common in the plains from May to early November, then disappear (Macdonald 1906, Smythies 1986); one was seen in Thailand in January ( Jacob and Martron 1996); in Sri Lanka they occur through the year (Legge 1880, Harrison 1999). In India they disappear early in the dry season. Indian birds pass across the Arabian Sea and appear in Oman and southern Arabia in migration

Jacobin Cuckoo Clamator jacobinus 323 (Gallagher and Woodcock 1980, Jennings 1981, Cornwallis and Porter 1982, Bryan 1988, Goodman and Jennings 1988, Ericksen 1994, 1995, Porter et al. 1996); they sometimes occur from April through July and breed there. Indian Jacobins also move across the Indian Ocean and migrants appear in the Seychelles (Phillips 1984). In East Africa, wintering Indian birds are indistinguishable in size and color from northern African birds. In Tanzania where breeding African Jacobins are scarce the birds are silent, in molt and not breeding from December to April (Moreau 1937a), the season when Indian Jacobins are there. C. j. jacobinus breed in summer in India mainly south of 15°N and in Sri Lanka and occur there in winter (Ali and Ripley 1969, Harrison 1999), although there are more winter records in Africa (Clancey 1960, Friedmann 1964a, Irwin 1988). Occasional in Zimbabwe (Irwin 1981, Tree 1983); and I collected a male east of Lilongwe, Malawi, on 27 March 1967 (UMMZ 216698), by its small size (wing 139), fresh plumage and small gonads a bird from southern India.

Habitat and general habits In open woodland and scrub (Acacia, Combretum, Terminalia, mopane Colophospermum mopane), dry grasslands with scattered trees, thorny jungle, thickets; in India mainly lowlands, plains and hills to 2000 m, in migration to 4200 m in Himalayas, in Africa mainly below 1500 m, occurs to 3000 m. Feed mainly in trees and bushes, hopping from branch to branch, peering under leaves and in forks of branches, and on the ground where they hop in search of food.After it catches a caterpillar, the bird works the prey back and forth through the bill and swings it from side to side until it is gutted and flattened, then swallows the insect. Common in much of the breeding range.

Food Insects, mainly hairy and smooth caterpillars, also grasshoppers, mantids, flying termites, moths, beetles; forest snails. Cuckoos take eggs of host birds. A laying female with an egg nearly ready to lay in her oviduct had eggshell and yolk in her crop, another

had eggshell and near-hatch embryo of a songbird (bulbul Pycnonotus sp.) in her crop, and a breeding male had yolk on his bill (Payne 1973a, UMMZ). They take berries and fruits (Gaston 1981, Roberts 1991, Ali 1996).

Displays and breeding behavior Territorial, socially monogamous. Early in the breeding season, the cuckoos are conspicuous in noisy courtship chases in flight and in calling from trees. Later in the season, pairs remain together much of the day and call without display before they copulate. In courtship feeding, the male flies from the female while she remains perched then slowly raises and lowers her tail, her crest held erect then lowered. She calls, the male brings her a hairy caterpillar after he wipes off the hairs, and she accepts it before the pair copulates (Godfrey 1939, Liversidge 1971, Gaston 1976). The male leads the host from the nest while the female enters the nest and lays (Skead 1951, Friedmann 1956, Liversidge 1971). A cuckoo mobbed by a Paradise Flycatcher Terpsiphone viridis in the Eastern Cape, South Africa, was a breeding male (UMMZ 213025)—I did not see the female at the nest.The female lays on alternate days, in the early morning between 0700 and 0900 hrs, then ovulates her next egg and copulates later in the morning (Liversidge 1961, 1971). She lays in sets of 3–4 eggs with a few days between sets, on average 19–25 eggs in a season (Payne 1973a, 1974). In southern Africa, black birds are common in the Eastern Cape and white-bellied birds are common from Transvaal northwards. The two color phases occur in equal numbers in southern Natal (West et al. 1964), where I observed them near Estcourt in December 1966. In four pairs both birds were black and in four other pairs both were white-bellied, suggesting a degree of assortative mating (Payne 1973a).Their calls were the same. In the Eastern Cape where white-bellied birds are uncommon (10%), they pair with black birds. The male often associates with the female and appears to draw the host bird away from the nest, while the female cuckoo enters the nest to lay (Liversidge 1971).

324 Jacobin Cuckoo Clamator jacobinus

Breeding In the rains, in N India with the SW monsoon from June to August or September (Ali 1946, Gaston 1981), in the S Nilgiris from January to March (Whistler 1928, 1935), in Sri Lanka mainly from February to May, with some from November to June and August (Whistler 1944). In the Arabian peninsula a female had “ovaries well developed, one almost perfectly formed egg” on 22 May (1901–1902) (AMNH 624521), a pair was taken together on 26 June 1946 (BMNH), and birds sing and display in summer in North Yemen (Brooks et al. 1987). In West Africa they breed seasonally from May to July (Bannerman 1951, Lamarche 1980, Morel and Morel 1990), in Ethiopia March to October (Urban and Brown 1971), in Somalia in May (Ash and Miskell 1998), in E Africa with the local rains, east of the rift valley from March to May, in the rift and westward from March to August (Brown and Britton 1980), in Angola in November (Heinrich 1958, Dean 2000), in southern Africa from October to March (Friedmann 1948, 1964a, Benson and Benson 1977, Irwin 1981, 1988, Maclean 1993, Skinner 1996, Dean 2000); in Natal they lay from October to December (Dean 1971, Payne 1973a). Brood-parasitic. Hosts, in Pakistan they parasitize Common Babbler Turdoides caudatus (Roberts 1991), in India their hosts are babblers (Common Babbler, Jungle Babbler T. striatus, Large Grey Babbler T. malcolmi and Yellow-billed (Whiteheaded) Babbler T. affinis) (Legge 1880, Baker 1906, 1907, 1934, Ali and Whistler 1937, Ali and Ripley 1969, Gaston 1976, Johnsingh and Paramanandham 1982); in Mauritania, Sudan, Kenya and Ethiopia their hosts are babblers (Fulvous Chatterer Turdoides fulvus, Rufous Chatterer T. rubiginosa) (Erlanger 1905, Friedmann 1948, Serle and Morel 1977, Huels 1982, Nikolaus 1987, Morel and Morel 1990), in southern Africa the hosts are bulbuls Pycnonotus (Black-eyed Bulbul P. barbatus, Cape Bulbul P. capensis, the latter feeds fruit to the nestling cuckoos, and Sombre Bulbul Andropadus importunus), Fiscal Shrike Lanius collaris and Paradise Flycatcher Terpsiphone viridis and sometimes other birds (Friedmann 1956, 1964a, Payne and Payne

1967, Liversidge 1971, Irwin 1988, Skead 1995, Tarboton 2001). Eggs are plain blue in India and Sri Lanka, usually plain blue in Africa north of 14°S, rarely white in Mali and Kenya, and all plain white in southern Africa. Eggs are 24.5 ⫻ 19.3 mm (C. j. jacobinus) in India (Baker 1906, 1908a, 1942), 26.5 ⫻ 21.9 mm (C. j. serratus) in Africa (Harrison 1971, Irwin 1988). In Asia, the eggs look like those of the babbler host but are rounder, paler blue and without gloss; in southern Africa the eggs are dull white (unlike any host eggs). The female cuckoo takes a host egg from the nest in some populations (Liversidge 1971, Payne 1973a, Gaston 1976, Rowan 1983). The incubation period is 11–12 days. Nestlings beg upwards in response to movement of the nest or to a call, but after day two they respond only when the nest is touched. They beg from a crouched position and do not move the wings until day 11, when the wings are used more in movement into or out of the nest than in begging. Nestlings grow from 6–7 g at hatching to 12 g by day 3, 24 g by day 6, 36 g by day 9, 48 g by day 10 to about 56 g before they fledge. They reach 50% of fledging weight by day 7. Young cuckoos fledge in 14–15 days in South Africa; young that left at 11 days were early and unable to fly (Liversidge 1971).Young are ready to leave the nest in 11–12 days in India (Gaston 1976, Johnsingh and Paramanandham 1982); they leave prematurely when they are disturbed (MacDonald 1957). Fledged cuckoos are fed by their foster parents for 2–3 weeks after leaving the nest and are fully grown except for a short tail by day 26. Nestling Jacobin Cuckoos do not evict the host eggs from a nest, although they often crowd out the host young or monopolize the parental care and food in the brood, depending on the body size of the host young. In South Africa no eggs were found under a nest where a cuckoo chick had hatched, and when tested with a finger on the back the cuckoo chicks did not assume an eviction posture like that of Red-chested Cuckoo Cuculus solitarius (Liversidge 1971).Young cuckoo chicks spread their wings and feet over the host eggs or host young in a nest, and this posture prevented the young hosts from begging when the parent came in to feed.

Little Cuckoo Coccycua minuta 325 Young bulbul chicks appeared out of the nest, dead, from the second to fourth day after hatching. It is unknown whether the cuckoo chick pushed them out from the nest or the host young were starved and weakened in the nest and were removed by their parents, although the posture of the cuckoo chick in the nest and its interference with the begging behavior of the host chick are likely to have led to the loss of the young bulbul. Sometimes two cuckoos fledge in a brood (Liversidge 1971), and a pair of Fiscal Shrikes once reared three cuckoos (West et al. 1964). Host young infrequently fledge with young cuckoos in South Africa, but often fledge with them in India where the cuckoos are smaller and the hosts are communally-breeding babblers with several adults to feed the brood (Liversidge 1971, Gaston 1976, Johnsingh and Paramanandham 1982). In South Africa, Cape Bulbuls Pycnonotus capensis are small, on average 38 g, whereas in India, Jungle Babblers Turdoides striatus are 66 g and Common Babblers T. caudatus are 38 g. The cooperative care of their brood by parents and helpers in the cooperatively breeding groups of babblers and the large size of the Jungle Babbler may account for the greater success of Indian babblers than African bulbuls, in terms of

host young surviving to fledge from a nest with a nestling cuckoo. Regionally the proportion of nests that are parasitized is high for the main host species, in southern Africa 16% of Cape Bulbul nests, 13% of Sombre Bulbul nests and 12% of Black-eyed Bulbul nests (Tarboton 2001). Of 50 eggs laid during four years, 10 produced fledged cuckoos, with success of the cuckoo ranging from 0 to 35% from year to year. Cuckoo parasitism affects the breeding success of their hosts. In South African Cape Bulbul where 36% of nests in one local population were parasitized (Liversidge 1971), 24% of the parasitized nests fledged a young bulbul, in contrast to 33% success of unparasitized nests. In only one of 10 bulbul nests that fledged a cuckoo did a bulbul also fledge. In Kenya, a cuckoo and babblers fledged together in a brood (Huels 1982). In northern India, in 38 Jungle Babbler nests, 71% were parasitized, breeding success of unparasitized nests was 59% and of parasitized nests 32%; in 31 Common Babbler nests, 31% were parasitized, host breeding success of unparasitized nests was 77% and of parasitized nests 33%. In this area one or more babblers fledged in 30% of the nests where cuckoos fledged (Gaston 1976).

Genus Coccycua Lesson, 1830 Coccycua Lesson, 1830, Traité d’Ornithologie, ou tableau méthodique des ordres, sous-ordres, familees, tribus, genres, sous-genres et races d’oiseaux. Levrault, Paris, pp. 1659, livre 2, 142. Small New World cuckoos with a rounded to graduated tail. The type, by monotypy, is Coccyzus minutus Vieillot (Peters

1940), which has rufous plumage and short rounded wings.Wing shape is rounded in Coccycua minuta, intermediate in C. pumila and pointed in C. cinerea, although not as pointed as in most Coccyzus cuckoos.Three species.

Little Cuckoo Coccycua minuta (Vieillot, 1817) Coccyzus minutusVieillot, 1817, Nouveau Dictionnaire d’Histoire naturelle . . ., 8, p. 275. Polytypic. Two subspecies. Coccycua minuta minuta (Vieillot, 1817); Coccycua minuta gracilis (Cabanis and Heine, 1863). Other names: Piaya minuta (Vieillot, 1817).

Description ADULT: Sexes alike, crown cinnamon brown sometimes raised in a bushy crest, back rufous brown, wing rufous brown, tail rufous black with terminal white spots; underparts, throat and breast rufous chestnut, belly dark buffy gray, under tail coverts

326 Little Cuckoo Coccycua minuta blackish, under wing coverts rufous to brownish buff, underside of flight feathers cinnamon with the tips brownish slate; eye-ring red, iris red, bill greenish or grayish yellow with dusky tip, feet dull blue to grayish black. JUVENILE: Above dark gray brown washed rufous, wing dark rufous, tail blackish with no white on tip (some birds have a narrow reddish-brown tip), tail narrower than in adult; underparts dark sooty gray with a rufous wash, belly and under tail coverts blackish; iris brown, bill dark to black, feet black. NESTLING: Undescribed. SOURCES: AMNH, BMNH, CM, FMNH, LSU, MCZ, MVZ, ROM, UMMZ, USNM, ZSM.

Subspecies Plumage varies geographically in brightness and intensity of color. Birds in northern South America are dark, notably in Trinidad and the Orinoco and Amazon river deltas. Birds of western Ecuador are pale, as are birds in Panamá and some birds in Bolivia. Several color variants have been described as subspecies.There are two main regional variants in plumage, dark and pale, and in geographically intermediate regions the birds are intermediate between these plumage extremes. Coccycua minuta minuta (Vieillot, 1817); upperparts dark purplish brown, throat and breast nearly as dark as upperparts, breast rufous buff, belly dark gray; Colombia and Ecuador east of the Andes, Venezuela and the Guianas south to Amazonian Peru and Brazil and SC Brazil (Goiás, Mato Grosso),Trinidad; Coccycua minuta gracilis (Cabanis and Heine, 1863); upperparts squirrel-rufous, throat and breast pale rufous, belly light gray; E Panamá, Colombia north and west of the eastern Andes, western Ecuador and perhaps Bolivia. The other subspecies that have been described are variants on these plumage color themes and the populations do not differ consistently between regions or are intermediate: (1) Coccycua minuta panamensis (Todd, 1912); described with throat and breast paler than C. m. minuta, of E Panamá and N

Colombia west of Gulf of Urabá is a synonym of C. m. gracilis. Todd indicated that birds of Panamá were paler than birds of Ecuador, but a larger series shows no difference. Coccycua minuta chaparensis (Cherrie, 1916) of N Bolivia was said to be darker than panamensis, but specimens from Bolivia (Todos Santos and Rio Beni) in FMNH and LSU are as pale or paler than birds in Panamá; they are intermediate in breast color between the paler western C. m. minuta and darker eastern C. m. gracilis. (3) Coccycua minuta barinensis (Aveledo Hostos and Peréz Chinchilla, 1994) of western Venezuela, are intermediate in color between gracilis and minuta and do not consistently differ from C. m. minuta which are variable but not consistently so across geographic regions, as noted also by Zimmer (1930). (4) Birds of the Orinoco delta were described as C. m orinocensis by Cherrie (1916), but these are not consistently darker than C. m. minuta elsewhere in northern South America, noted also by Cory (1919). Birds of the lower Amazon are also variable in color (Pinto 1947).

Measurements and weights C. m. minuta, Guyana, Suriname, French Guinea and NE Brazil to Pará and Mato Grosso: Wing, M (n ⫽ 13) 101–107 (105.2 ⫾ 2.4), F (n ⫽ 5) 104–108 (105.0 ⫾ 2.0); tail, M 138–158 (147.9 ⫾ 6.4), F 146–163 (156.2 ⫾ 6.9); bill, M 13.8–23.5 (20.1 ⫾ 2.1), F 16.8–22 (18.8 ⫾ 3.2); tarsus, M 20.5–27.9 (23.8 ⫾ 2.2), F 22–26.4 (24.1 ⫾ 1.7) (AMNH, ANSP, FMNH, UMMZ); C. m. gracilis, Ecuador: Wing, M (n ⫽ 5) 102– 109 (104.6 ⫾ 3.6), F (n ⫽ 7) 97–108 (103.4 ⫾ 4.3); tail, M 148–170 (150.3), F 133–170 (144.1 ⫾ 12.2) (AMNH, LSU, UMMZ). Weight, M (n ⫽ 11) 26–54 (40.9), F (n ⫽ 16) 30.9–50 (40.1) (ANSP, FMNH, LSU, UMMZ, Haverschmidt and Mees 1994). Wing formula, P6 ⬎ 7 ⬎ 5 ⬎ 4 ⬎ 8 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 25 cm. Small cuckoo like a half-size Squirrel Cuckoo Piaya cayana, rufous above, rufous to dark gray below; raises and lowers the tail slowly.

Little Cuckoo Coccycua minuta 327

Voice Calls, a coy laughing nasal “kyaaah hoo-hoo-hoohoo” or “tyoooooo aannh anhh-anhy-anhh”, the first note a long drawn-out call starting at 0.9 kHz and rising to 1.1 near the end, the later notes starting around 0.6 kHz and rising to a loud peak at 1.0–1.1 kHz before dropping again, the first two “aanhh” notes higher and the last two falling in pitch, the first phrase lasting 0.7 sec and the “aannh” notes each 0.1 sec, the entire phrase about 1.2 sec and distinguished by kazoo-like harmonics at 2 and 3 times the lower pitch. Another call is a “tchek-wrraaawk” with the first note a sharp rising call and the second a crow-like rising buzzy note lasting about 0.6 sec (Hardy et al. 1990). Other calls reported are harsh clucks “tchek” like those of a tree frog, a sharp “quienk”, and a hoarse weak “geep” (Hilty and Brown 1986).

Range and status Northern South America east of the Andes from Colombia and the Guianas south to Brazil (Amazonas, Pará, Amapá, Mato Grosso, Goiás), Ecuador and Peru, and Bolivia (a pair seen in the southeastern tip on the eastern bank of the Río Negro, Dpto. Santa Cruz, on the border with Paraguay); Trinidad; Panamá and Darién. In Colombia they occur near Lago de Sonso,Valle, and are not seen on the Pacific coast (Phelps and Phelps 1958, Wetmore 1968, Haffer 1975, Hilty and Brown 1986, Remsen and Traylor 1989, Sick 1993, Haverschmidt and Mees 1994, AOU 1998, Capper

et al. 2001). Resident. Uncommon to fairly common and local in distribution.

Habitat and general habits Tropical lowland evergreen forest edge, secondary forest, clearings, mature mangroves, river-edge forest, typically in dense, low vegetation; thickets, dense canebrakes, vines, shrubby growth, pastures, often near water, mainly coastal and along major rivers in the Guianas, in flooded várzea forest, in savanna uplands in Amazonian Brazil, in Cecropia forest on seasonally flooded islands in Amazonian Peru and oxbow lakes in lowlands of Ecuador. Occur from sea level to 600 m, on east slope of the Andes a few follow river valleys up to 800–900 m, on the west slope locally to 1500 m in Ecuador near Mindo and Maquipucuna, and they occur from lowlands to 1600 m in Colombia and to 2000 m in Venezuela, and in riverine scrub in Bolivia. Uncommon birds in the understory of dense shrubby woodlands, they forage in understory to mid-story levels, and are not birds of primary forest (Wetmore 1968, Dubs 1982, Terborgh et al. 1984, Hilty and Brown 1986,Tostain et al. 1992, Stotz et al. 1996, 1997, da Silva et al. 1997, Clay et al. 1998, Capper et al. 2001, Ridgely and Greenfield 2001; ANSP, FMNH, LSU).

Food Insects including caterpillars, wasps, beetles, bugs and flies, also spiders and tarantulas (Pinto 1953, Wetmore 1968, Hilty and Brown 1986,Tostain et al. 1992, Haverschmidt and Mees 1994, Novaes and Cuhna Lima 1998; ANSP, LSU, UMMZ).

Breeding In Trinidad they nest in July (Belcher and Smooker 1936), in Colombia birds are in breeding condition from January to May (Hilty and Brown 1986), in Suriname eggs occur from June to September (Hellebrekers 1942, Haverschmidt and Mees 1994), in French Guiana a young bird fledged from its nest in early March ( Tostain et al. 1992), in Brazil near Belém a young was attended to by an adult in March (Willis 1983a), and in Bolivia females had large ova in July (LSU 132010) and November (FMNH 334287).The nest is an open mass of small

328 Dwarf Cuckoo Coccycua pumila twigs, with a shallow cup, built in a dense shrub or tree. Eggs are dull white, 24 ⫻ 19 mm, weight 0.4 g, clutch 2 (Belcher and Smooker 1936, Hellebrekers

1942, Schönwetter 1964, Haverschmidt and Mees 1994, Novaes and Cuhna Lima 1998).The incubation and nestling periods are unknown.

Dwarf Cuckoo Coccycua pumila (Strickland, 1852) Coccyzus pumilus Strickland, 1852, Jardine’s Contributions to Ornithology, p. 28. [“Trinidad” ⫽ Venezuela] Monotypic.

Weight, U (n ⫽ 6) 36 (Ralph 1975), M 45.7 (LSU), F 39.1 (FMNH). Wing formula, P7 ⫽ 6 ⬎ 8 ⫽ 5 ⬎ 4 ⫽ 9 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Description

Field characters

ADULT: Sexes alike, crown pale slate gray, back brownish gray, wing brownish gray, darker near the tip, tail short, brownish gray with subterminal band black and narrow band of white on tip; underparts, face, throat and upper breast rufous, belly and under tail coverts white or buff, under wing coverts dark buff, inner edge of primaries buff; eye-ring red (extends to lores and above eye), iris red, bill black, feet black.

Overall length 21 cm. Small gray cuckoo with a rufous throat and breast, and a short, square to rounded tail.

JUVENILE: Face, throat and upper breast pearly gray; eye-ring yellow, iris brown. NESTLING: At hatching, skin is blackish to dark gray tinted yellowish green, nearly naked with a few tan hair-like feathers on dorsal tracts. The mouth lining is red, the palate has three small whitish spots in front, a very small spot one on each side of the palate cleft, and a large whitish lateral spot raised in relief, with pointed papillae on each side of the palate cleft, and two white transverse bars behind these. SOURCES: AMNH, BMNH, CM, CU, DMNH, FMNH, LSU, MCZ, MVZ, USNM.

Measurements and weights Wing, M (n ⫽ 17) 92–112 (102.2 ⫾ 5.5), F (n ⫽ 10) 97–110 (102.3 ⫾ 4.2); tail, M 95–111 (105.2 ⫾ 7.9), F 94–122 (106.0 ⫾ 9.2); bill, M 16.2–23 (18.9 ⫾ 1.8), F 16.9–21 (19.1 ⫾ 1.7); tarsus, M 17.8–24.5 (20.6 ⫾ 2.1), F 17.5–23.3 (20.3 ⫾ 2.2) (AMNH, CM, CU, DMNH, FMNH, USNM).

Voice Calls, “churr,” grating “trrr trrr trrr . . .”, a wideband chatter with the peak amplitude in each note falling from 4 to 2.8 kHz, the notes each 0.2 sec long and repeated 2 per sec for several seconds; also a dawn song “kööa kööa” (Ralph 1975, Hilty and Brown 1986, Hardy et al. 1990).

Range and status Semi-arid regions in South America, one in northern South America in Colombia (Pacific coast near Buenaventura, Caribbean region from Atlántico east to western base of Santa Marta Mts, north end of Perijá Mts, middle Magdalena Valley), and one east of the Andes from Norte de Santander to Arauca, Meta,

Dwarf Cuckoo Coccycua pumila 329 Carimagua and W Florencia) and Venezuela (Zulia, Táchira east to Anzoátegui, Delta Amacuro and E Monagas, southern Apure, NW Bolívar, Amazonas south to Río Ventuari, and Margarita I); also in northern Brazil (Ilha Maracá, Roraima) (Short 1975, Meyer de Schauensee and Phelps 1978, Hilty and Brown 1986, Fjeldså and Krabbe 1990, Sick 1993, Whittaker 1995). “Trinidad” as the source of the holotype is thought to be an error (Benson 1999); and a record at Concepcion in Entre Ríos,Argentina (Sclater and Hudson 1889), is singular. Resident in some areas, migratory in others. In Colombia, individually marked birds remain on their territory all year; in Venezuelan savannas, the birds appear in season with the rains (Phelps and Phelps 1958, Ralph 1975, Thomas 1978, 1979, Hilty and Brown 1986, Whittaker 1995). Fairly common.The species range has expanded where forests have been cleared for pastures (Ralph 1975).

Habitat and general habits Dry lowland scrub, deciduous forest, gallery forest, secondary forest, open woodland, pastures with scattered trees, also at western edge of Amazon rainforest (not breeding); sea level to 1000 m, and to 2600 m near Bogotá, Colombia (Ralph 1975, Thomas 1978, 1979, de la Peña 1986, Fjeldså and Krabbe 1990, Hilty and Brown 1986, Sick 1993, Stotz et al. 1996). The cuckoos live and feed in trees, shrubs and on the ground, usually 1–6 m above ground, slowly peering around in the inner and outer branches, and dropping to the ground near cover (Ralph 1975, Hilty and Brown 1986, Whittaker 1995). A breeding pair uses an area of about 10 ha, and neighboring pairs often share their foraging areas (Ralph 1975).

Food Insects, caterpillars and treehopper (membracid) nymphs (Ralph 1975).

Displays and breeding behavior Territorial, defend the area around the nest. Monogamous, although one female mated with two males and both males helped rear the young. Pair maintains social contact with “churr” calls, and adult on the nest gives this call. Male courts female bring-

ing an insect to her, before the pair has a nest; the birds give a sputter-churr as they approach each other, the female crouches slightly, quivering the wings, the male hops onto the female’s back allowing her to grasp the other end of the caterpillar. Pair maintains cloacal contact for as long as 25 sec, then male pumps and rotates his tail and slides off to one side, maintaining cloacal contact a few more seconds, then leaves female with the caterpillar. Male and female build the nest. Male brings insects to female while she is incubating and brooding the young, as often as three times in an hour. Pair stays together and nest more than once in a season (Ralph 1975).

Breeding Near Cali, Colombia, nests are found from January through September (Ralph 1975), and in Venezuela the birds have large gonads from March through August (POS).The nest is a flimsy platform of twigs, tendrils, vines and leaves, built 1–7 m above ground in the canopy of a short tree, the nest 10–15 cm in diameter with a cup of 2–3 cm, sometimes lined with a dry leaf or two. The eggs are dull white, 24.9 ⫻ 19.6 mm, clutch 2–3(4) (Ralph 1975, de la Peña 1986, USNM). Eggs are laid as far apart as three days. Two females sometimes lay in the same nest, when two eggs appear on the same day. Incubation period is 12 days and incubation begins with the first egg; the young hatch asynchronously often a day apart. Both sexes incubate and care for the young, with incubation sessions of 30–156 min and brooding sessions of 20–25 min.The male incubates and broods at night. Nestlings are fed c. 3 times an hour. The young nestling quivers its wings in begging. By day 4 pinfeathers appear on the wings, tail and body, and the nestling can grasp the twigs of the nest with its feet. On day five the eyes are almost fully open. On day six some feathers erupt from the sheaths, and the nestling can stand and defecate over the edge of the nest. The nestling fledges in 11–12 days and remains in the nest tree a few more days. Begging call of nestlings and fledglings is a low grating continuous “eeeeeee” heard at close range, and when young leave the nest area they give a loud, short “churr”. Nest success is 56%, and predation accounts for most (57%) of the nest mortality (Ralph 1975).

330 Ash-colored Cuckoo Coccycua cinerea

Ash-colored Cuckoo Coccycua cinerea Vieillot, 1817 Coccyzus cinereus Vieillot, 1817, Nouveau Dictionnaire d’Histoire naturelle . . ., 8, p. 272. (Paraguay) Monotypic.

Description ADULT: Sexes alike, crown gray, back and rump brownish gray, wing brownish gray, tail rounded, brown above and light gray below,T1 to T3 brown gray with a narrow white tip and T4 to T5 with a black subterminal band and a narrow white tip; underparts, throat and breast light gray, belly white, flanks and under tail coverts buff, bend of wing white, wing lining pale buff; eye-ring red, iris red to brown, bill black with base of mandible silver, feet gray. JUVENILE: Crown and back brown with a rufous wash, tail brown, narrow and pointed at tip, no distinct subterminal band and no white tip; underparts, throat and breast grayish white, breast gray, center of belly white, flanks, sides of belly and under tail coverts buff. NESTLING: Hatchling, skin undescribed; mouth with white papillae, 5 on the palate and 2 pairs around the pharynx (Wetmore 1926). SOURCES: AMNH, ANSP, BMNH, CM, FMNH, LSU, MSNM, UMMZ, USNM,YPM.

Measurements and weights Wing, M (n ⫽ 11) 97–111 (106.8 ⫾ 3.8), F (n ⫽ 9) 104–116 (109.6 ⫾ 4.3); tail, M 98–114 (106.0 ⫾ 5.4), F 105–114 (109.4 ⫾ 4.2); bill, M 17.6–23.0 (20.4 ⫾ 2.2), F 17.5–24 (21.7 ⫾ 2.2); tarsus, M 17.5–23.3 (20.9 ⫾ 1.9), F 19–23 (21.0 ⫾ 1.5) (AMNH, LSU, MSNM, UMMZ). Weight, M (n ⫽ 3) 34.8–39.0 (37.1), F (n ⫽ 5) 43–57 (45.4) (FMNH, LSU, Belton 1984, Contreras 1979). Wing formula, P8 ⫽ 7 ⬎ 6 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 25 cm. Small gray cuckoo with short wings, pale gray throat and breast and a rounded, not graduated tail.

Voice A plaintive, petulant “cow-w cow-w cow cow”, or “che-rro, che-rro, che-rro”. The notes are drawnout descending whistles, starting at 1.0 and ending at 0.8 kHz, each note about 0.3 sec and repeated at 2 notes per sec for several seconds (Wetmore 1926, Hilty and Brown 1986, Hardy et al. 1990, de la Peña and Rumboll 1998).

Range and status South America, breeding in Brazil (Mato Grosso do Sul (Rio Ivinheima), São Paulo, Rio Grande do Sul), Paraguay, Uruguay and northern Argentina (Salta, Entre Rios, Corrientes, Tucumán, Buenos Aires) (Laubmann 1939, Cuello and Gerzenstein 1962, Gore and Gepp 1978, Contreras 1979, Short 1975, 1976, Belton 1984,Willis and Oniki 1985, Narosky and Di Giacomo 1993, Sick 1993, Straube and Borschein 1995, de la Peña and Rumboll 1998, Maurício and Dias 2000, YPM). In their breeding range in the southern temperate region, they are seasonal in spring and summer from late October to April (Belton 1984,Willis and Oniki 1985, Maurício and Dias 2000).The northern extent of their breeding range and winter range is not well known.They occur as nonbreeding visitors during the austral winter from March to October in subtropical and tropical Brazil (Amazonia, Rondônia, Bahia, Goiás) (Pinto 1938, Stotz et al. 1996, 1997, Sick 1997) and are not known to breed in the tropical region. In Paraguay they occur in winter in June, September and October in molt and with small gonads, and in the austral summer they breed there, with records of occurrence in Dptos. Bóqueron, Presidente Hayes and Ñeembucú (Hayes 1995, AMNH, UMMZ). Reported once at Leticia, Amazonas, Colombia, in July (Hilty and Brown 1986). Accidental or in

Squirrel Cuckoo Piaya cayana 331

Habitat and general habits ?

Scrub woodland, chaco scrub, cerrado woodland, savannas, tropical deciduous forest, riverine forest (Short 1975, Straube and Borschein 1995), in winter in low forest (Stotz et al. 1997); lowlands to 900 m. Flight is direct and low, birds revealing the short wings and long tail (Wetmore 1926).

Food Insects (LSU).

Breeding passage in SE Peru with records at Tambopata Reserve and Pampas del Heath (Parker 1982, Clements and Shany 2001), passage migrant in northern Bolivia (Gyldenstolpe 1945b, Remsen and Traylor 1989), and in southern Bolivia there are specimens during the austral breeding season, from 20°S at Lagunillas on 15 Nov 1936 (ANSP 143581) to 18°S at Dpto. Santa Cruz on 1 Jan 1989 (LSU 150536). Uncommon on breeding grounds in temperate and subtropical southern South America, and scarce in tropical regions.

Near Porto Allegre, Brazil, a female had a shelled egg in the oviduct in November (Belton 1984, AMNH 812993), and in Paraguay birds have large gonads in December and January (Short 1976).The nest is a platform of sticks and leaves, lined with leaves, placed in a bush or tree (Wetmore 1926, Short 1975, Sick 1993). Eggs are white, 25.5 ⫻ 19 mm (Schönwetter 1964, de la Peña 1986). Clutch size, 2–3 in Paraguay (Short 1975), 3(4) in Argentina (de la Peña 1986). Incubation and nestling periods are unknown.

Genus Piaya Lesson, 1830 Piaya Lesson, 1830, Traité d’Ornithologie, ou tableau méthodique des ordres, sous-ordres, familees, tribus, genres, sous-genres et races d’oiseaux. Levrault, Paris, pp. 1659, livre 2, 139. Type, by original designation, Cuculus cayanus Gmelin i.e. Linnaeus.The origin of the word

is an indigenous name “piaye” meaning “the Devil’s minister” referring to some perceived sinister quality of this furtive bird (Mitchell 1957).The genus is characterized by large size, rufous plumage, short wings, a short bill and a long graduated tail.Two species.

Squirrel Cuckoo Piaya cayana (Linnaeus, 1766) Cuculus cayanus Linnaeus, 1766, Systema Naturae (ed. 12), 1, 170. Other common names: Alma de gato Polytypic. Fourteen subspecies. Piaya cayana cayana (Linnaeus, 1766); Piaya cayana mexicana Swainson, 1827; Piaya cayana macroura Gambel, 1849; Piaya cayana circe Bonaparte, 1850; Piaya

cayana mehleri Bonaparte, 1850; Piaya cayana thermophila Sclater, 1860; Piaya cayana nigricrissa (Cabanis, 1862); Piaya cayana mesura (Cabanis and Heine, 1863); Piaya cayana pallescens (Cabanis and Heine, 1863); Piaya cayana cabanisi J.A.Allen, 1893; Piaya cayana insulana Hellmayr, 1906; Piaya cayana obscura Snethlage, 1908; Piaya cayana mogenseni

332 Squirrel Cuckoo Piaya cayana (Peters, 1926); Piaya cayana hellmayri (Pinto, 1938).

Description ADULT: Sexes alike, upperparts rich chestnut, tail long and graduated, rufous above with black subterminal band and white tip, face rufous; underparts, throat and upper breast rufous, lower breast and belly gray, under tail coverts black, tail dark chestnut with broad white tips below; bare skin around eye yellowish green (red east of the Andes), iris red, bill yellowish green, legs and feet light bluish gray. JUVENILE: Plumage similar, more lax and fluffy than adult, tail without black band, tail feathers narrower and with more pointed tips, central pair with terminal and subterminal marks indistinct; skin around eye gray, iris brown, bill gray horn, feet gray. NESTLING: Skin black, covered with hair-like down; mouth lining red (black in adults) (Skutch 1983). SOURCES: AMNH, BMNH, CM, FMNH, LSU, MCZ, MVZ, NMW, RMNH, ROM, UMMZ, USNM, ZSM.

History and subspecies Piaya cayana thermophila Sclater, 1860; dark rufous chestnut above, belly and under tail coverts dark gray to black, tail black below with narrow white tips; Mexico from E San Luis Potosí and S Tamaulipas south to Veracruz, Oaxaca and Isthmus of Tehuantepec on Atlantic region to Pacific region in Oaxaca from Chiapas west to near Tehuantepec City, through Central America to Panamá and NW Colombia; Piaya cayana mexicana Swainson, 1827; paler rufous chestnut above, tail rufous with black subterminal marks and broad white tips below; Mexico on Pacific slope and inland south to Isthmus of Tehuantepec; Piaya cayana nigricrissa (Cabanis, 1862); plumage darker than thermophila, belly and under tail coverts blackish; W Colombia to slopes of C Andes, W Ecuador, NW and E-C Peru;

Piaya cayana mehleri Bonaparte, 1850; plumage more rufous than mexicana, lighter throat and breast grading to light gray on belly, tail rufous below; NE Colombia east of Gulf of Urabá to Magdalena Valley and along W slope of eastern Andes, east along coast of Venezuela to Paria Peninsula; Piaya cayana mesura (Cabanis and Heine, 1863); plumage like nigricrissa, smaller, overlap in size; Colombia east of eastern slope of E Andes, eastern Ecuador; Piaya cayana circe Bonaparte, 1850; plumage above slightly more rufous than mehleri and paler than cayana;Venezuela south of Lake Maracaibo; Piaya cayana cayana (Linnaeus, 1766); belly ashy gray, under tail coverts darker gray, color like thermophila except belly and under tail coverts not as dark, tail black with white tips below; northern South America from E and S Venezuela through the Guianas and south to Brazil and north bank of the lower Amazon; Piaya cayana insulana Hellmayr, 1906; like cayana but under tail coverts black;Trinidad; Piaya cayana obscura Snethlage, 1908 (includes Piaya cayana boliviana Stone 1908); darker above (less rufous), under tail coverts black; central Brazil south of Amazon from Rio Juruá east to the Tapajóz, south to upper Gy-Paraná, Chanchamayo Valley in E-C Peru southeast into northern Bolivia; Piaya cayana hellmayri Pinto, 1938; upperparts more tan (less rufous); Brazil south of the Amazon from Santarem to mouth of Amazon, coast of northern Maranhão; Piaya cayana pallescens (Cabanis and Heine, 1863); upperparts paler, more tan and less red or gray; E Brazil; Piaya cayana cabanisi J. A. Allen, 1893; more rufous, throat rusty, belly and under tail coverts pale; Brazil (Central Mato Grosso and Goiás to 21°S, north of São Paulo); Piaya cayana macroura Gambel, 1849; upperparts dark, belly and under tail coverts blackish, long purple tail; SE Brazil (Bahia) to Paraguay, NE Argentina and Uruguay; Piaya cayana mogenseni Peters, 1926; upperparts lighter rufous, throat rusty, underparts dark gray not blackish; S Bolivia (Santa Cruz) to NW Argentina.

Squirrel Cuckoo Piaya cayana 333 The regional variation and subspecies of Squirrel Cuckoos have been described earlier by Stone (1908), Ridgway (1916), Cory (1918), Hellmayr (1929), Naumburg (1930), Junge (1937b), Pinto (1938, 1947), Griscom and Greenway (1941), Gyldenstolpe (1945a) and Haffer (1975). A few intermediates are known between P. c. thermophila and P. c. cayana which occur on opposite slopes of the cordillera (Binford 1989). Additional geographic forms have been described, including the birds of Sonora, Sinaloa and Colima, Mexico, described as a form P. c. extima van Rossem 1930, paler than P. c. mexicana further south (Davis 1960); this geographic variation in plumage color is subtle. The geographic variation among regions suggests limited dispersal between forested regions in the Neotropics.

Measurements and weights Piaya cayana thermophila, Mexico (Campeche, Yucatan, Chiapas): Wing, M (n ⫽ 12) 141–157 (146.1 ⫾ 4.7), F (n ⫽ 12) 138–152 (145.7 ⫾ 4.2); tail, M 246–284 (270⫾12.8), F 245–272 (263.2 ⫾ 12.1); bill, M 23.5–28.4 (26.6 ⫾ 1.4), F 23.4–27.8 (25.4 ⫾ 1.4); tarsus, M 33.8–38.8 (35.4 ⫾ 1.7), F 33.9–39.1 (36.0 ⫾ 1.6) (UMMZ); Panamá: Wing, M (n ⫽ 10) 138.2–147.8 (142.5), F (n ⫽ 10) 133.1– 145.8 (140.0); tail, M 237–279 (257), F 238–277 (253); bill 30.5–34.1 (32.1), F 30.5–33.5 (31.6); tarsus, M 33.7–37.5 (36.2), F 34.1–37.4 (35.8) (Wetmore 1968). Weight, M (n ⫽ 33) 73.0–137.0 (104.0), F (n ⫽ 28) 76.0–129.4 (100.3) (MVZ, UMMZ). Wing formula, P7 ⫽ 6 ⫽ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 46 cm. Large and slender cuckoo, mostly rufous, with a very long, graduated, whitetipped tail. Often perches in a tree with tail hanging down, or runs along branches with tail extended, like a squirrel.

Voice Calls, (1) a repeated “wheep-wheep-wheep . . .”, each “wheep” note rising from 1.8 kHz to peak

amplitude at 2.8 kHz then falling, the note 0.2 sec long and repeated in a series at 2 notes per sec; (2) a two-note “chick-kaw” with the first note rising and the second falling, both with rich harmonics at 1, 2, 3 and 4 kHz, the call lasting 0.3 sec with a few seconds between calls; (3) a three-noted dry rolling “hic-a-ro” or “geep-kareer”, with the first note “wheep” followed by a drawn-out crow-like drawl with rich overtones, the phrase lasting 0.8 sec, the “wolf-whistle”; (4) a dry rattle “churrrr” and (5) a sharp “kip!” (Hilty and Brown 1986, Stiles and Skutch 1989, Hardy et al. 1990, Howell and Webb 1995, de la Peña and Rumboll 1998, Royte 2001).

Range and status Mexico and Central America south through South America west of the Andes from Colombia to NW Peru, and east of the Andes from Colombia through Venezuela and the Guianas to southern Brazil, Uruguay and NE Argentina (Cuello and Gerzenstein 1962, Haffer 1975, Wetmore 1968, Hilty and Brown 1986, Remsen and Traylor 1989, Stiles and Skutch 1989, Fjeldså and Krabbe 1990,Tostain et al. 1992, Narosky and di Giacomo 1993, Sick 1993, Haverschmidt and Mees 1994, Howell and Webb 1995, AOU 1998, de la Peña and Rumboll 1998). Resident, largely absent from offshore islands. Fairly common in liana-covered trees and along forest edges. Often in pairs. Common, c. 4 pairs /100 ha in floodplain várzea forest (Terborgh et al. 1984, 1990, Robinson and Terborgh 1997).

334 Squirrel Cuckoo Piaya cayana

Habitat and general habits Tropical lowland evergreen forest, tropical deciduous forest, old secondary forest, pine-oak woodland, gallery and river bluff forest, mangroves, flooded tropical evergreen várzea forest, generally more in open forest than closed forest.They live in the canopy in open forest, tangles of vine, coffee plantations, shrubs, pastures with scattered trees, watercourses in dry regions, high ground forest, also in residential areas in Brazil; they avoid dense lowland forests but occur in forest edge and treefall gaps. Occur from lowlands at sea level to 1500 m and occasionally up to 2700 m (Miller 1963, Hubbard and Crossin 1974, Short 1975, Thomas 1978,Willis and Oniki 1991, Belton 1984, Wetmore 1986, Hilty and Brown 1986, Stiles and Skutch 1989, Fjeldså and Krabbe 1990,Tostain et al. 1992, Sick 1993, Stotz et al. 1996, 1997, Cohn-Haft et al. 1997, Bencke and Kindel 1999, Ridgely and Greenfield 2001). Arboreal gleaning insectivores, they keep under cover, hop along limbs or through vines in bounding leaps, forage by concealment in foliage, scan the leaves, then pounce on an insect. Solitary, usually in pairs, sometimes in mixed flocks, they follow army ant swarms in the forest interior (Dickey and Van Rossem 1938, Sutton 1951, Oniki and Willis 1972,Willis and Eisenmann 1979,Willis 1983, Jullien and Thiollay 1998). Birds move quietly in outer canopy in leaves, run along branches, and leap from branch to branch.They catch flying ants in flight. Squirrel Cuckoos are capable of sustained flight, glide from tree to tree across open spaces with bursts of shallow wingbeats, and use their long tails as rudders in flight (Wetmore 1968, Haverschmidt and Mees 1994).

Food Insects, mainly caterpillars (green or hairy), grasshoppers, walking sticks, moths, beetles, ants, bees, wasps (Dickey and Van Rossem 1938); tree

frogs (Cintra and Sanaiotti 1990) and some fruit (Blake and Hanson 1942).

Displays and breeding behavior Socially monogamous. One bird brings sticks and leaves to the nest, and the other builds the nest. A pair nests repeatedly in a season, and a breeding adult feeds its mate (Skutch 1966, 1983, Oniki and Willis 1999).

Breeding In Sonora, Mexico, early May to June (Russell and Monson 1998), in Jalisco during the rains in July (Selander and Giller 1959, Rowley 1984), in Oaxaca in May (Binford 1989), in Yucatan in July (UMMZ), in Costa Rica in both dry and wet seasons from January to August (Skutch 1983, Stiles and Skutch 1989), in Colombia from April to July, copulation also observed in January (Hilty and Brown 1986), in Trinidad in July (Brooker and Smooker 1936), in Suriname from March to July and September (Hellebrekers 1942), in French Guiana from April to September (Tostain et al. 1992), in Pará, Brazil, in March, May and July (Stone 1929, Pinto 1953). Nest is a shallow platform of leaves, or sticks and green leaves, or rootlets, placed in vines, thick shrubbery or a tree, often more than four meters above ground, where it looks like hanging rubbish in the lianas (Rowley 1984). Eggs are chalky white, becoming stained brown, 35 ⫻ 26 mm, and weight of a fresh egg is 0.74 g. The clutch is 2–3. Two females sometimes lay in a nest (Sick 1993). Incubation is 18 days. Both parents incubate, each in a bout of 6–8 hours before being relieved by the mate. Both parents care for the young bringing large insects about once an hour to the brood. The young leave the nest in eight days or so after hatching, before they can fly (Hellebrekers 1942, Skutch 1966, 1983, Wetmore 1968, de la Peña 1986).

Black-bellied Cuckoo Piaya melanogaster 335

Black-bellied Cuckoo Piaya melanogaster (Vieillot, 1817) Cuculus melanogaster Vieillot, 1817, Nouveau Dictionnaire d’Histoire naturelle . . ., 8, p. 236. [Cayenne]. Monotypic.

Description ADULT: Sexes alike, crown light gray, back rufous chestnut, wing rufous with outer vanes P7-P10 brown-black and the primaries with blackish tips, tail rufous, rectrices with an intergrading subterminal black band and a broad white tip; underparts, chin and throat buff, breast rufous, belly and under tail coverts black; bare orbital skin pale blue around eye, large yellow spot in front of eye and greenish at rear corner of eye, iris dark brown to red, bill red, feet lead gray to black. JUVENILE: Similar to adult, but tail feathers narrow, with short white tips (8–12 mm, vs 12–14 in adult). NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, LSU, MCZ, ROM, UMMZ, USNM, ZFMK, ZMB.

Measurements and weights Wing, M (n ⫽ 12) 131–142 (135.8 ⫾ 5.2), F (n ⫽ 10) 129–141 (135.4 ⫾ 3.7); tail, M 193–239 (221.1 ⫾ 12.6), F 193–242 (218.5 ⫾ 17.1); bill, M 30.0–35.5 (32.9 ⫾ 1.7), F 27.7–35.5 (33.6 ⫾ 2.3); tarsus, M 30–37.4 (33.6 ⫾ 2.3), F 29–37 (32.2 ⫾ 2.2) (AMNH, FMNH, LSU, UMMZ). Weight, M (n ⫽ 6) 86.0–111.5 (98.8), F (n ⫽ 3) 78.7–109 (93.4) (Haverschmidt and Mees 1994; FMNH, LSU, USNM, ZFMK). Wing formula, P7 ⫽ 6 ⫽ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 8 ⬎ 9 ⬎ 10.

Field characters Overall length 38 cm. Large cuckoo with a red bill and gray cap, bright face colors, rufous above, belly black, long tail.

Voice Calls, a loud clear two-note phrase “huweep-hew”, the first note rising from 1.6 to 2.2 kHz then falling rapidly to 1.5 kHz, the second note softer and rising to 1.7 kHz then falling to 1.2 kHz, the phrase lasting 0.3 sec and repeated once a second for as long as a minute, or sometimes with the second note itself drawn out to 0.3 sec and falling from 2.4 to 1.8 kHz; and (2) a scratchy rattling squeeze-toy “hu-jijijiji-yaaaah”, the first note short rising then falling, the second phrase a rising rattle, the third a descending note, the entire phrase lasting nearly a second and repeated after a few seconds (Hilty and Brown 1986, Hardy et al. 1990, Ridgely and Greenfield 2001).

Range and status South America from Colombia (east of Andes, Macarena Mts), Venezuela (Amazonas, Bolívar), Guyana (Ituribisi, Kamakabra and Mazaruni Rivers, Bartica, Wineperu, Kamakusa, Rockstone, Tumatumari, Merumé Mts), Suriname, French Guinea, eastern Ecuador and eastern Peru, NW and NE Bolivia, and Brazil (Amazonia and Pará south to northern Mato Grosso). Resident. They occur together with the more numerous Squirrel Cuckoo P. cayana in Ecuador but are absent in parts of Bolivia where Squirrel Cuckoos occur; in Amazonian non-flooded tierra firme forest they are

336 Dark-billed Cuckoo Coccyzus melacoryphus more common than Squirrel Cuckoos (Cohn-Haft et al. 1997). Uncommon lowland forest birds, they are easily overlooked. They are sensitive to habitat disturbance with loss of their forests (Chubb 1916, Griscom and Greenway 1941, Gyldenstolpe 1945a, Phelps and Phelps 1958, Snyder 1966, Hilty and Brown 1986, Parker and Remsen 1987, Haverschmidt and Mees 1994, Sick 1993, Stotz et al. 1996, Ridgely and Greenfield 2001).

et al. 2001, J.V. Remsen, in litt.). Lowlands to 800 m, in Ecuador mainly below 400 m, once at 900 m north of Lumbaquí (Snyder 1966, Hilty and Brown 1986, Remsen and Traylor 1989, Sick 1993, Haverschmidt and Mees 1994, Ridgely and Greenfield 2001). Feed in forest canopy, solitary and in mixed-species foraging flocks (Cohn-Haft et al. 1997).

Habitat and general habits

Insects, including grasshoppers, katydids, caterpillars, beetles, hemiptera, homoptera, flies and ants, and spiders and Chilopoda (Haverschmidt and Mees 1994, AMNH, CM, FMNH, LSU, UMMZ, USNM).

Tropical lowland evergreen forest, upland forest, sand-belt forest, less often in shrubby growth away from forest and in savanna woodland. In Guyana and Suriname, P. melanogaster is absent near the coast and in savanna, it occurs in forests of the interior. In Amazonia, P. melanogaster is restricted to upland tierra firme forest where P. cayana tends to be a várzea riverine / transitional forest species; they occur together where their ranges and habitats come together, while south of the range of C. melanogaster, P. cayana occurs into tierra firme forest (Cohn-Haft et al. 1997, Stotz et al. 1997, Borges

Food

Breeding A bird was in breeding condition in the upper Orinoco in April, and an adult fed a young bird in French Guiana in July (Hilty and Brown 1986, Tostain et al. 1992). The nest is undescribed. Eggs are white, 30 ⫻ 23 mm (Schönwetter 1964). Incubation and nestling periods are unknown.

Genus Coccyzus Vieillot, 1816 Coccyzus Vieillot, 1816, Analyse d’une nouvelle Ornithologie élémentaire , p. 28. Arboreal cuckoos of the New World, with the tail medium to long and graduated in shape.Type, by monotypy,“Coucou de Caroline” Buffon ⫽ Cuculus americanus Linnaeus (Peters 1940). The name derives from the Gr. kokkuzo, to cry cuckoo (kokkux, the cuckoo). The genus includes small cuckoos with pointed wings and a graduated tail. Although most Coccyzus are small and have long, pointed wings, the group also includes the large, short-winged birds formerly recognized in two other genera in the West Indies:

Saurothera Vieillot, 1818, type, by monotypy, “Coucou à long bec” Buffon ⫽ Cuculus Vetula Linnaeus; and Hyetornis Sclater, 1862, type, by monotypy, Cuculus pluvialis Gmelin 1788. These large island cuckoos are derived from smaller continental Coccyzus cuckoos and are included in the genus Coccyzus.The smaller lizard-cuckoos C. vieilloti, C. longirostris and C. vetula are nearly identical in songs but are distinctive in plumage and are retained as species. Seven species of small to medium-sized Coccyzus plus four in Saurothera and two in Hyetornis, total 13 species in Coccyzus.

Dark-billed Cuckoo Coccyzus melacoryphus Vieillot, 1817 Coccyzus melacoryphus Vieillot, 1817, Nouveau Dictionnaire d’Histoire naturelle . . ., 8, p. 271. (Paraguay)

Monotypic. Other common names: Azara’s Cuckoo.

Dark-billed Cuckoo Coccyzus melacoryphus 337

Description

Voice

ADULT: Sexes alike, crown and nape gray, back grayish brown, wing dark gray-brown, tail long, graduated, above bronzy brown, below black with broad white tips (T1 brown with black tip, T2 blackish with narrow white tip), face with black mask from in front of eye through eye to ear; underparts, narrow gray band down side of neck, chin and throat, breast to under tail coverts buff, under wing coverts whitish to buff; eye-ring gray or olive yellow, iris dark brown, bill black, feet slate gray to olive.

Song, six to eight resonant, buzzy kazoo-like notes, “zneaah-znaah-znaah-znaao-znaao-znau”, each note with two peak amplitudes, one at 0.8 kHz and the other at 1.0 kHz and with rich overtones, the note lasting 0.2–0.3 sec and repeated 3 per sec in a series of about 8 notes, the series rising then falling slightly in pitch. Other calls include a dry rattle “dddddrr”, “charr-charr-chao-chur-churr” or “cherrrrooouuuu” (Belton 1984, Hilty and Brown 1986, Hardy et al. 1990, de la Peña and Rumboll 1998). Nestling begging call, a hatchling “peep” (Friedmann 1927).

JUVENILE: Similar to adult, duller, crown and nape brown, wing sometimes rufous and coverts with indistinct buff tips, tail feathers narrow, pointed and with indistinct gray tips. NESTLING: At hatching skin, dusky orange, nearly naked with white hair-like down, and skin over eye dusky green; the gape swollen and white (Friedmann 1927). SOURCES: AMNH, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM, ZMA, ZMUC, ZSM.

Measurements and weights Paraguay: Wing, M (n ⫽ 9) 113–123 (117.9 ⫾ 3.2), F (n ⫽ 11) 115–123 (119.2 ⫾ 2.6); tail, M 126–148 (135.6 ⫾ 7.5), F 134–144 (137.2 ⫾ 5.0); bill, M 21.2–24 (22.3 ⫾ 0.8), F 20.2–24.5 (22.6 ⫾ 1.5); tarsus, M 22.8–26.1 (24.3 ⫾ 0.9), F 23.2–25.2 (24.0 ⫾ 0.6) (UMMZ). Weight, M (n ⫽ 14) 42–54 (45.6), F (n ⫽ 7) 50–66.5 (54.2) (AMNH, ANSP, FMNH, UMMZ, Belton 1984, Haverschmidt and Mees 1994, Magalhães 1999). Wing formula, P8 ⫽ 7 ⬎ 6 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 27 cm. Cuckoo with gray brown upperparts, gray crown and rich buff underparts, black mask through eye and cheek, and a long tail with large white tips. The wing is gray brown, not rufous.

Range and status South America: Colombia, Venezuela and the Guianas, Trinidad, Ecuador, west-central and northern Peru east of the Andes, Brazil (Amazonas, Pará, Bahia, Minas Gerais, São Paulo, Rio Grande do Sul), Bolivia, Paraguay, Uruguay and northern and central Argentina, and the Galapagos Archipelago. Resident in some areas, migratory in others. In Uruguay and Argentina they occur during the austral summer from October to January (Hudson 1920, Wetmore 1926, Friedmann 1927, Pinto 1938, Laubmann 1939, Mitchell 1957, Cuello and Gerzenstein 1962, Belton 1984, de la Peña 1986, Dubs 1992, Sick 1993, Hayes et al. 1994, Narosky and di Giacomo 1993, de la Peña and Rumboll 1998). In Paraguay and Bolivia most records are in passage migration from September to December and in April; some birds remain during the austral winter (Laubmann 1939, Hayes et al. 1994,

338 Dark-billed Cuckoo Coccyzus melacoryphus Jahn et al. 2002, CM). In southern Amazonian Brazil they are nonbreeding visitors mostly in the austral winter (Zimmer et al. 1997). In northern Brazil birds occur in Pará from April to August (Novaes and Cunha Lima 1998, CM) and at Manaus and Tonantins, Amazonas, from May to September or October, perhaps mainly as nonbreeders (M. Cohn-Haft, in litt.; CM). In northern South America, they occur and breed in the boreal summer, in the Guianas from May to September (Tostain et al. 1992, Haverschmidt and Mees 1994). Both breeding and nonbreeding birds may occur in the austral winter in the llanos of Venezuela (Phelps and Phelps 1958,Thomas 1978, 1979) although all specimens (FMNH, POC, USNM) were taken between May and October. In Colombia they occur year-round; most specimen records are from April to October and others in January, February, and March (Hilty and Brown 1986, ANSP, CM, FMNH, MVZ, USNM). In eastern Ecuador nonbreeding birds occur from late April to early October (as early as February and as late as November), while a resident breeding population occurs in the lowlands and foothills of the southwest (Ridgely and Greenfield 2001). In the Galapagos, where this cuckoo is the only land bird not peculiar to the Archipelago, the birds appear in all months, with local appearances on islands where they had not been seen, when weather and winds disperse the birds from other islands (Swarth 1931, Curry and Stoleson 1988). One record in northern Chile (Arica, Johnson 1972), accidental on the Falkland Islands (Bennett 1937), the West Indies (Grenada: Schwartz and Klinikowski 1965), Texas, Panamá and Clipperton I (Stager 1964, AOU 1998). Fairly common.

Habitat and general habits Tropical deciduous forest, tierra firme forest, gallery forest, second growth forest, river-edge forest, shade trees of coffee plantations, tropical lowland evergreen forest; humid woodland edges, thickets and mangroves, drier scrub, forest openings, transitional forests with seasonal flooding, lake margins, marshy areas, thick vine growth, river margins, coastal plains on the dry Galapagos in scrub of cacti Opuntia, small trees, spiny shrubs and salt bush

Cryptocarpus. Mainly in lowlands, from sea level to 1000 m, they occur to 2100–2400 m in Colombia, to 2800 m in semi-arid montane valleys of Ecuador and the eastern Andean slopes of Peru, they are regular to 2500 m in the dry valleys of Cochabamba and Chuquisaca in Bolivia, and they occur to 1800 m in NW Argentina (Wetmore 1926, Short 1975, Belton 1984,Terborgh et al. 1984, de la Peña 1986, Ervin 1989, Remsen and Traylor 1989, Tostain et al. 1992, Haverschmidt and Mees 1994, Stotz et al. 1996, Cohn-Haft et al. 1997). Migrants appear at high altitudes, one at 3600 m in Cuzco, Peru, in late November and December (Fjeldså and Krabbe 1986). These cuckoos are birds of forest edge and tree-fall openings, and generally solitary in behavior (Cohn-Haft et al. 1997).

Food Insects, mainly caterpillars (hawkmoth larvae) and grasshoppers, also stick insects, beetles and ants (Pelzeln 1871, Friedmann 1927, Belcher and Smooker 1936, Ervin 1989, Haverschmidt and Mees 1994).

Displays and breeding behavior The male brings an insect to the female in courtship (Ralph 1975).

Breeding In Colombia they nest in October (Hilty and Brown 1986), in Venezuela they nest and have large gonads from May to August (Thomas 1979, POC), in Suriname from May to August (Haverschmidt and Mees 1994), in southern Brazil in November (Belton 1984), in the chaco of Bolivia in December, and in Argentina from October to January (Friedmann 1927, Eisentraut 1935, de la Peña 1983, 1986, 1995). In the Galapagos, cuckoos appear and breed from January to March, and to May in seasons of heavy rains (ENSO, El NiñoSouthern Oscillation) (Harris 1982, Curry and Stoleson 1988, Ervin 1989). The nest is a flat platform of sticks and twigs, 8–15 cm across and 4–8 cm deep, built in trees and bushes, often under a leafy canopy, placed 1–3 m above ground or in mangroves at water’s edge. Nest

Yellow-billed Cuckoo Coccyzus americanus 339 material is also brought after eggs are laid and during incubation. Eggs are green to pale blue or bluish-white, 28 ⫻ 21 to 30 ⫻ 23 mm (Friedmann 1927, Eisentraut 1935, Belcher and Smooker 1936, Hellebrekers 1942, Schönwetter 1964, de la Peña 1983, 1986, Ervin 1989). Clutch size is 2–3 in Suriname, 2 in Brazil, 2–5 in Argentina, 4–5(6) in the Galapagos (Hudson 1920, Friedmann 1927, Snethlage 1928, Belton 1984, de la Peña 1983, 1986, 1995, Ervin 1989, Haverschmidt and Mees 1994). Two females sometimes lay in a nest

(Stresemann 1934, Sick 1993). Incubation period is 11–12 days. The eggs sometimes hatch synchronously (Hudson 1874), or more often they hatch asynchronously over several days. Nestlings open their eyes on day 3, and the young fledge in 7–9 days (Dabbene 1926, Ervin 1989, de la Peña 1995). The breeding cycle resembles that of other nesting cuckoos, with a short incubation period, asynchronous hatching, rapid growth of chicks, and fledglings that move about through the vegetation before they can fly.

Yellow-billed Cuckoo Coccyzus americanus (Linnaeus, 1758) Cuculus americanus Linnaeus, 1758, Systema Naturae (ed. 10), 1, p. 110. (Carolina ⫽ South Carolina) Monotypic.

Description ADULT: Sexes alike, above, unmarked olive brown lightly glossed with bronze, wing primaries rufous on inner web, tip darker brown, secondaries S1-S2 rufous with darker brown tip, others olive brown, tail long, graduated, blackish with large white tips below (T1 olive-brown and without white tips,T2 blackish with narrow white tip,T3 to T5 with large white tips longer than the width of the feather,T5 outer edge white nearly to the base); underparts white from throat to under tail coverts, under wing coverts buffy white; eye-ring yellow to gray, iris dark brown, bill black above, yellow below. Eye-ring color may change with age; breeding birds on the nest have either yellow or gray eye-ring (photos, UMMZ). JUVENILE: Plumage similar to adult, tail with rectrices narrow and tapered near tip, on T3 to T5 whitish tail spots large, indistinct and grading into the proximal vane, white outer edge on T5; underparts white from chin to under tail coverts; eye-ring gray to pale yellow, iris dark brown, bill black above and gray below. NESTLING: Skin blackish, at hatching nearly naked with a few dusky gray hair-like down feathers, frontal apterium paler gray than bill or neighboring

skin; corner of gape unswollen and red, tomia white, mouth lining red, palate red with five white papillate patches (the lateral patches larger than the medial patches; the anteromedial patch sometimes divided by the narrow rostral slit of the choana), two pairs of large elongate mediolateral patches along the pars caudalis of the choana, tongue red with a pair of small white papillated patches on the alae linguae, a large white patch of papillae on the dorsal surface and a black tip, feet blue-gray (Nolan 1975, UMMZ, RBP). SOURCES: AMNH, BMNH, CM, FMNH, LSU, MCZ, MVZ, POC, ROM, TCWC, UMMZ, UNM, USNM, ZMA.

Measurements and weights Michigan: Wing, M (n ⫽ 11) 126–142 (138.8 ⫾ 5.8), F (n ⫽ 12) 139–150 (144.2 ⫾ 3.6); tail, M 138–156 (145.5 ⫾ 5.9), F 132–157 (147.3 ⫾ 7.5); bill, M 22.5–28.7 (25.6 ⫾ 1.7), F 23.2–27.5 (25.5 ⫾ 1.1); tarsus, M 22.3–25.4 (23.9 ⫾ 0.8), F 23.4–25.7 (24.4 ⫾ 0.9) (UMMZ); Texas: Wing, M (n ⫽ 10) 135–148 (140.9 ⫾ 3.7), F (n ⫽ 10) 131–152 (144.7 ⫾ 6.4); tail, M 128–147 (138.4 ⫾ 6.6), F 132–152 (142.8 ⫾ 6.4); bill, M 23.7–26.6 (24.1 ⫾ 1.6), F 23.9–27.2 (25.6 ⫾ 0.9); tarsus, M 21.2–25.8 (23.7 ⫾ 1.5), F 22.1–26.7 (24.0 ⫾ 1.3) (TCWC); Arizona: Wing, M (n ⫽ 12) 138–146 (142.4 ⫾ 3.9), F (n ⫽ 7) 137–156 (149.9 ⫾ 6.0); tail, M 147–157 (149.7 ⫾ 3.4), F 150–156 (153.1 ⫾ 2.0);

340 Yellow-billed Cuckoo Coccyzus americanus bill, M 26.9–29.0 (27.9 ⫾ 1.0), F 26.8–31.3 (28.9 ⫾ 1.7); tarsus, M 22.2–27.4 (24.3 ⫾ 1.6), F 25.0–27.7 (25.9 ⫾ 1.0) (UMMZ). Weight, Michigan, breeding season: M (n ⫽ 11) 42.8–70.9 (58.3) (before 12 September), F (n ⫽ 10) 53.6–78.0 (65.3) (excludes laying females and all females after 12 Sept.), (most females larger than 70.0 had an ovum larger than 10 mm or an egg in the oviduct, or were in migration and had much body fat) (UMMZ). Weights are as heavy as 110 g in fat birds before migration (UMMZ 157809), and as light as 31g after a long over-water migration (Voous 1983). Texas, breeding season: M (n ⫽ 2) 53.5–57 (55.3), F (n ⫽ 5) 61–74 (67.5) (TCWC). Wing formula, P8 ⬎ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10. Birds in western North America are larger on average and were described as a subspecies C. a. occidentalis (Ridgway 1887). Because size (wing, tail) overlaps considerably between populations in the east and west of the breeding range, no subspecies are recognized here. Birds on their wintering grounds cannot be traced to eastern or western breeding populations by their size, and there are no banding recoveries to indicate wintering areas of birds breeding in eastern or western populations. A molecular genetics study that reported differences between eastern and western populations (Pruett et al. 2001) was based on a small sample (eight birds in their breeding range) and the difference in base sequence in a highly variable gene was less than 1%. In a larger sample (66 birds), no fixed differences were found in base sequences between eastern and western birds (Fleischer 2001).

Field characters Overall length 30 cm. Slender cuckoo, brown above and white below, with a long graduated tail, large white spots on tips of tail feathers, white edge to outer tail feathers, wing rufous in flight, and bill black above and yellow below. Juveniles differ from Black-billed Cuckoo C. erythropthalmus in uniformly colored wing coverts (buff edges in Blackbilled Cuckoo); flight feathers of the wing are rufous in both species, but with less dark brown in Yellow-billed Cuckoo. Juveniles differ from Mangrove Cuckoo C. minor in lacking whitish bars

on the upperparts (barred in Mangrove Cuckoo). The nestlings differ in the color of natal down, gray in Yellow-billed Cuckoo and white in Black-billed Cuckoo.

Voice In breeding season, a rapid throaty “kikikikikiki-kaka-ka-ka-ka-ka”, broad-band calls with the first notes abrupt and given about 10 per sec, the later notes slower, loud and guttural, broad-band descending notes that fall from 2.0 kHz to 1.2 kHz, last 0.1 sec and are repeated about 5 per sec. Sometimes give a sharp knocking call “kikikikiki” alone, and a soft repeated rain-call “coo” or “kowlp”, a plaintive call that lasts about 0.3 sec and drops in pitch from 0.9 to 0.6 kHz, often with a catch near the end and the note repeated twice per second for several seconds. Female gives soft “coo” calls in courtship and male brings nest material to her. In “coo” call the adult inflates the throat and perches in a crouched posture.Adult gives soft rapid “kikiki” when approaching the nest with food, often in groups of 3 short notes at a rate of 10 per sec, with a pause of 0.3 sec between groups. Begging call of young, a “buzz” appears by day 1 and grows louder over a few days, the older nestlings give “cuckcuk-cuk-cur-rr-r-rrr” with shorter calls after they are fed, and fledged birds give a “croak” (Preble 1956, Potter 1980, Hamilton and Hamilton 1965, Hardy et al. 1990, Hughes 1999a).

Range and status North America, mainly in the eastern United States from the Great Lakes region, New York and southern New England south to Florida and Texas, and in extreme southern Canada from Manitoba south of Lake Winnipeg, southern Ontario and Quebec (Godfrey 1986, Gauthier and Aubry 1996). Scarce and local in the interior western United States, where they breed along the Snake River, Colorado River, Platt River and Great Salt Lake; they are rare in the Central Valley of California and in southern Arizona (Oberholser 1974, Banks 1988a, 1990, Franzreb and Laymon 1993, Pyle 1997, Hughes 1999a)). In the Chihuahuan desert region of the United States and Mexico they occur along rivers and in the Guadalupe, Davis, Chisos and Carmen

Yellow-billed Cuckoo Coccyzus americanus 341

mountains (Van Tyne and Sutton 1937, Miller 1955,Wauer and Ligon 1977). In northern Mexico they occur from Chihuahua, Coahuila, Nuevo León, and Caribbean lowlands in Tamaulipas and Yucatan, inland on the N Pacific slope, the Cape Region of Baja California, and they lived on the Colorado River delta in the early 20th century when water still flowed south from the United States into Mexico (AOU 1998). July sightings in Veracruz, Guatemala and El Salvador suggest breeding there (Andrle 1967, Wendelken and Martin 1986,Thurber et al. 1987, Howell and Webb 1995). In the northern West Indies they breed in Cuba, Hispaniola and Puerto Rico, rarely in Jamaica and the Virgin Islands (Gundlach 1874, Wetmore 1927, Paynter 1955, Edwards 1957, Klaas 1968, Howell and Webb 1995, Raffaele et al. 1998). Migratory, they move at night and they orient by the pattern of stars (Hamilton and Hamilton 1965). Over-water flight in autumn in the east follows the Florida panhandle as indicated by fall-outs at lighthouses and by the time series of birds appearing in Hispaniola and Puerto Rico, or over the Gulf of Mexico in flight of 2000–3000 km covers a long distance of 2000–3000 km to the West Indies from

where the birds move further on, or it covers 4000 km in direct flight from their breeding range to mainland South America (Crawford and Stevenson 1984, Mc Nair et al. 2002).They migrate over land in southern Mexico from September and October and in April (Binford 1989, Howell and Webb 1995), in Guatemala from September to December (Land 1970), in El Salvador from August to October (less common in spring, in May) (Thurber et al. 1987), in Costa Rica from August to November (less common in spring,April to June) (Stiles and Skutch 1989), and in Panamá from August to November, and in spring in April and May (Willis and Eisenmann 1979, Wetmore 1968). Seasonal in Florida, where they are rare in winter, northbound migrants appear through May into June and southbound migrants by July; and by early September they appear on islands where they do not breed (Fisk 1979, Stevenson and Anderson 1994). In the southern Bahama Islands, they are occasional in summer (no breeding records) and common in autumn migration in October (Buden 1987). In the Cayman Islands they appear in migration from late April to June and in autumn from October to November (rarely as early as August) and from late April to June, with one winter record in January (Bradley 2000). In the summer rainfall areas of Sonora, cuckoos arrive in late June and wait a month for rains before they breed (Russell and Monson 1998). In Central America, cuckoos are mainly passage migrants. In some years a few appear in winter (Guanacaste, Costa Rica, (Stiles and Skutch 1989); Barro Colorado I, Panamá, (Willis and Eisenmann 1979)); nearly all move on to South America for winter, perhaps because of dry conditions in winter in Central America (Willis 1980). Cuckoos are breeding residents and passage migrants in the Greater Antilles and the Caribbean (Stockton de Dod 1981, Raffaele et al. 1998) with only one record in January (Kepler and Kepler 1978). In the Netherlands Antilles the main passage in autumn is from mid-October to early November (Voous 1983), with one record in Aruba in winter ( January, ZMA 28900). In northern South America the cuckoos are migrants on their way to winter further south. They occur in Colombia mainly in passage months and not in winter (December to February). In the

342 Yellow-billed Cuckoo Coccyzus americanus

?

eastern Caribbean they appear from early October to early November, in the Bocas Is between Venezuela and Trinidad in October and November, in the arid Caribbean zone of Colombia from September to December, and along the arid Caribbean coast of Venezuela they pass through in December (Russell 1980, Hilty and Brown 1986, Bosque and Lentino 1987, Hayes and Somad 2002, McNair et al. 2002, CM, FMNH, LSU, POC, USNM). Inland in Venezuela there are specimens mainly in passage months although with records in all months except December and January and only one in February (Zimmer and Hilty 1997, CM, FMNH, POC, USNM), and in the Guianas they occur in October, November and April (Haverschmidt and Mees 1994). In the central Amazon of Brazil they are reported in May and June (Stotz et al. 1992), a season suggesting birds in transit. Spring records of birds in passage are fewer in northern South America; the birds call as they pass the arid coast of Colombia and Venezuela from late April to mid-May (Russell 1980, Bosque and Lentino 1987). The birds winter from December to February in South America east of the Andes (Hilty and Brown 1986, Fjeldså and Krabbe 1990,Tostain et al. 1992, Haverschmidt and Mees 1994) south to Peru (FMNH), Bolivia (Remsen et al. 1986, Remsen and Traylor 1989, CM, FMNH, LSU), Brazil (south of the Amazon with most records in the east to Espiritu Santo and São Paulo, also in Mato Grosso and Piauí) (Naumburg 1930, Willis and Oniki 1985, Narosky and di Giacomo 1993, Sick 1993,

Parker and Goerck 1997, CM, LSU), Paraguay (Short 1976, Hayes 1995, FMNH, LSU, UMMZ), Uruguay (Wetmore 1926, Cuello and Gerzenstein 1962, Gore and Gepp 1978) and northern Argentina (Tucumán, La Rioja, Córdoba, Buenos Aires) (Meyer de Schauensee 1970, de la Peña and Rumboll 1998). In Ecuador and Peru they are rare winter residents (Ridgely and Greenfield 2001, LSU). Occasional cuckoos remain into early June in Suriname and Venezuela (Haverschmidt and Mees 1994, POC) and through summer in Colombia (Hilty and Brown 1986). Cuckoos are accidental as far north as SE Alaska (Ketchikan) and Newfoundland and west to Clipperton I (AOU 1998). Trans-Atlantic vagrant in autumn to Greenland, British Isles, Scandinavia, western Europe, Sicily, Morocco and the Azores, with more than 50 records in Europe (Cramp 1985, Iapichino and Massa 1989), and rarely seen in spring in Greenland (16 March 1874, ZMUC). Population densities on the breeding grounds vary across years, in apparent response to outbreak populations of caterpillars and cicadas. Common in North America from the Great Plains eastward, with centers of density in the southern Great Plains (Oklahoma, east Texas, Louisiana, also in Mississippi, Arkansas, Missouri, Tennessee and Alabama) (Price et al. 1995). In eastern North America the birds increased in numbers in the period 1965–1979, when there were widespread outbreaks of caterpillars (Robbins et al. 1986, Eastman 1991). Scarce in the west, where the species has disappeared along with the riparian woodlands in California, Arizona and southern Oregon (Gaines 1974). In Arizona, where habitat and census data was available along the Lower Colorado River from 1976 to 1986, cottonwood-willow habitat decreased by 31% and the cuckoos decreased by 16%; the number of cuckoos along the river in 1976 was about 846 pairs (Groschupf 1987). In California, the cuckoos sustained a major contraction in range and decline in numbers in the twentieth century (Gaines and Laymon 1984); the population declined to about 30 pairs with the loss of riparian breeding habitat and the widescale agricultural use of pesticides (Hughes 1999a). Numbers throughout North America have declined in recent years (31% in the

Yellow-billed Cuckoo Coccyzus americanus 343 period 1966–1993, 12% in 1984–1993) (Peterjohn et al. 1995, Price et al. 1995). Numbers in the wintering grounds in South America have decreased as well (Hughes 1999a). Local numbers can be high, as in Sonora where 40 birds were heard in one locality (Russell and Monson 1998), and in willowedged waterways in southern Oklahoma where five active nests were found in two hours one morning (Payne 1973a). Broadly sympatric with the Mangrove Cuckoo in Hispaniola, Puerto Rico and Jamaica, where they are present all year. A reported “hybrid” Yellow-billed Cuckoo ⫻ Black-billed Cuckoo (CM 149972, Parkes 1984) is a juvenile male Yellow-billed Cuckoo.The bird was found in Meridian, Butler County, Pennsylvania, on 22 Oct 1974.The rectrices are narrow and pointed, with indistinct whitish spots on T3–4, T5 is growing and has distinct white spots, the outer edge of T5 is white, the wing coverts are unmarked (wing coverts have whitish edges in juvenile C. erythropthalmus), the wing is rufous, not as bright as in the adults; and the wing formula is P8 ⬎ 7 (P8 ⫽ 7 in C. erythropthalmus). The crown is grayish brown as in C. americanus (darker brown in C. erythropthalmus).The wing is 142, the tail 134, the bill 22.5, the tarsus 23.0, all measurements within the ranges of male C. erythropthalmus and C. americanus. Bill color was not described in the fresh bird, but bill in the specimen is darker below than in adult C. americanus and is similar to other juvenile C. americanus.

Habitat and general habits Forest with clear dark understory, open woodland, shrubby pastures, riparian woodland, cottonwoods, willows, mesquite, thickets, tangles and swamps. In arid regions of the North American southern Great Plains, the southwest and California they seek out riparian woodlands and scrub habitat (Oberholser 1974, Short 1974, Rea 1983, Howell and Webb 1995). In migration they occur from tropical evergreen forest and deciduous forest to arid scrub (Raffaele et al. 1998), and migrants are seen to 4200 m in Venezuela and 2800 m in the Andes in Colombia and Ecuador (Hilty and Brown 1986, Fjeldså and Krabbe 1990). In winter in South America they live in woodlands (gallery forest, tropical deciduous forest, secondary forest), scrub and

bushy savanna, also in mangroves; generally in lowlands, and they often join mixed species foraging flocks (Terborgh et al. 1984, Hayes 1995, Robinson et al. 1995, Stotz et al. 1996). Skulking and secretive in behavior in dense vegetation, they forage in tree foliage, canopy and subcanopy, scan and search as they perch motionless, cock the head and wait until an insect moves, then hop to grab it.They wipe the caterpillar on a branch, shake it removing the guts, swallow tossing the insect into the air and catching it. And they snap at insects in flight, disturb insects when they brush them from foliage in flight through the canopy, and pursue prey on the ground. Territory size, a pair uses c. 20 ha in breeding season, and neighboring pairs overlap in feeding areas; cuckoos need woodland habitats of more than 40 ha (Gaines 1974, Hughes 1999a). The clearing of woodland habitats has led to the loss of most of the Yellow-billed Cuckoos in central California.

Food Large insects, mainly caterpillars and grasshoppers, and when these are present in outbreak proportions the cuckoos take tent caterpillars, fall webworms, gypsy moths, and periodic and annual cicadas. Other insects important in their diet include katydids, crickets, beetles, mayflies, dragonflies and sawfly larvae (Tenthredinidae). They also take phalangid harvestmen (Bendire 1895, Beal 1898, Wetmore 1916, Bent 1940, Witter and Kulman 1972, Jauvin 1996, Hughes 1999a), small frogs and Anolis lizards (Oberholser 1974,Voous 1983), eggs from bird nests and fruit (Rubus raspberries, elderberries, mulberries) in summer and autumn (Bendire 1895, Forbush 1927, Stevenson and Anderson 1994), and fruits and seeds in winter (Rappole et al. 1983). Food is similar to that taken by Black-billed Cuckoos. In open habitats,Yellow-billed Cuckoos take adult grasshoppers, although they prefer green, soft-bodied insects such as caterpillars and katydids when they feed their young (Laymon 1980, Potter 1980).

Displays and breeding behavior Male feeds the female in courtship (Bent 1940). Female solicits in crouch, head and tail raised, then lowers her tail to point straight down and raises it to point straight up, repeating the postures and

344 Yellow-billed Cuckoo Coccyzus americanus pumping action. Male approaches her, depresses his tail and moves it laterally toward exposed cloaca of female, grasps her with his bill, and the pair copulates for about 5 sec. He passes an insect to the female during copulation (Potter 1980) and she sometimes holds it in her bill for more than one copulation (Pistorius 1985). Male also brings a stick to female in courtship (Hendricks 1975), and he brings sticks while she is on the nest, which she accepts and adds to the nest (Eaton 1979).

Breeding In eastern North America the season is prolonged with time to rear two broods, in Ohio and Michigan from May to September (Trautman 1940, Wood 1951), in Oklahoma and Ontario from May to August (Sutton 1967, Peck and James 1983), in Maryland from May to July (Stewart and Robbins 1958), in Florida from April to August (Stevenson and Anderson 1994), in Texas from March to September (Oberholser 1974), in Cuba and Puerto Rico from May to July (Gundlach 1874, Wetmore 1916, Kepler and Kepler 1978), in Tamaulipas in May (Sutton et al. 1950) and in Sonora from June to August (Russell and Monson 1998). Nest is a platform of sticks, c. 28 ⫻ 36 cm with a shallow cup of 11 ⫻ 12 cm, lined with green and dead leaves, grass stems or other material or not lined at all, nest often saddled on a horizontal branch in a shrub or thorny tree, or willow over water, or in thick bush overgrown with wild grape and other vines, concealed, near ground level to 30 m in a shrub or tree. Nesting pair occasionally use the old nest of another bird. Nest lining is sometimes added after incubation begins. Eggs are light blue, smooth, not glossy, 30 ⫻ 23 mm. The clutch is 1–5 (in Ontario, Indiana, Kansas, Florida and Hispaniola usually 2–3; mean in Kansas 3.1, in Indiana 3.0, in Kentucky 2.5, in Cuba 4–5) (Wetmore 1926, Bent 1940, Mengel 1965, Sutton 1967, Johnston 1964, Nolan and Thompson 1975, Harrison 1978, Potter 1980, Hughes 1999a, Garrido and Kirkconnell 2000). Clutches are larger when insects are locally abundant. Occasionally more than one female lays in a nest: as many as eight eggs may appear in a nest. Female often lays eggs on alternate days (Payne 1973a, Sealy 2003). Both sexes incubate from the

first egg, the male and female taking turns in the daytime, the male at night (Potter 1980). Incubation period is 10–11 days (or as short as 9 days: Potter 1980); eggshell fragments are left in the nest. Both parents feed and brood the young and eat or remove their fecal sacs from the nest. Nestlings differ in size as a result of hatching on different days; the smallest often starve.The parents brood until the second-oldest young have feathers erupted from the sheaths. The nestling is partly covered with short pinfeathers on day 1 and can stand in the nest and flap its wings when it begs, its eyes open by day 3; it can grasp sticks with its feet by day 2; it stands on the nest rim, clutches the nest material and defecates over the rim by day 3–4, is active and covered with hair-like down on day 5, crouches low in the nest by days 4–5, the feather sheaths burst by days 6 and 7 (young are no longer brooded in daytime after day 6); the young remove the feather sheaths with the bill; all feather covers have broken open except those above the bill by day 8, the young fledge by days 8–9 (1–3 days later if weather is cool) when they are about 38 g. The total time from onset of incubation to fledging is 18–21 days (Audubon 1839, Bendire 1895, Forbush 1927, Roberts 1936, Bent 1940, Trautman 1940, Preble 1957, Hamilton and Hamilton 1965, Laymon 1980, Potter 1980, 1981, Hughes 1999a). Nestlings beg with the neck stretched upwards and the wings stretched and flapped to the side; they buzz when they beg. Yellow-billed Cuckoos and Black-billed Cuckoos sometimes lay in each other’s nests and in the nests of other birds, most frequently American Robin Turdus migratorius, Gray Catbird Dumetella carolinensis and Wood Thrush Hylocichla mustelina, and young cuckoos occasionally survive to fledge from the nest of another species (Bendire 1895, Nickell 1954b, Wiens 1965, Sutton 1967, Nolan and Thompson 1975, Peck and James 1983, Fleischer et al. 1985, Wolfe 1994, Hughes 1997a). In one case, two eggs appeared on the same day, apparently laid by two females (Sealy 2003). Breeding success, in California, 4 of 4 nests fledged young, although 3–5 eggs in a nest (mean, 3.5) produced only 1–2 fledged young (mean, 1.5) (Laymon 1980). In Indiana, in 7 nests there were 15 eggs but only one nest fledged any young (Nolan 1963).

Pearly-breasted Cuckoo Coccyzus euleri 345

Pearly-breasted Cuckoo Coccyzus euleri (Cabanis, 1873) Coccygus Euleri Cabanis, 1873, Journal für Ornithologie, 21, p. 72. (Cantagallo, Rio de Janeiro, Brazil). Also known as Coccyzus julieni Lawrence, 1864, but this name has been officially suppressed (Bulletin of Zoological Nomenclature 49, 178–9, 1992) (Banks 1988b,Willis and Oniki 1990). Monotypic.

Description ADULT: Sexes alike, upperparts gray brown lightly glossed with bronze, wing brown not rufous, inner webs of primaries white, tail long, graduated, blackish with large white tips (T1 brownish gray without spots,T2 to T5 brownish black, ventral narrow spots on T2 are 3 mm in length, broad spots on outer rectrices (T3 spots 13 mm, T4 spots 18 mm, T5 spots 22 mm), outer edge T5 white nearly to the base), face darker gray below and behind the eye; underparts from throat to breast pearly grayish white, belly and under tail coverts light pearly gray, center of belly white, under wing coverts pearly white; eye-ring gray to black, iris dark brown, bill black above, yellow or orange below and base of upper mandible yellow, feet gray to bluish gray. JUVENILE: Wing brown with fine pale edge, tail spots less distinct than in adult, whitish and grade into brown (spots lacking on T1,T2 spots are a terminal trace of whitish, T3 spots are 8 mm, T4 spots are 15 mm,T5 spots are 15 mm), rectrices narrower and more pointed than in adult (10 mm from tip,T4 is 9 mm wide in juvenile, vs 14–15 mm in adult), no dark face mark; iris, bill and feet as in the adult. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, CM, FMNH, MCZ, POC, RMNH, UMMZ, USNM, ZMB, ZMUC. Comments: Coccyzus euleri (and “C. julieni”) was long regarded as a synonym of C. americanus (Shelley 1891, Hellmayr 1913, Laubmann 1939, Gyldenstolpe 1945b), while others referred to euleri and julieni as a

subspecies of C. americanus and Ridgway thought the type of C. julieni was a young C. (americanus) euleri (Ridgway 1916). Later, Hellmayr (1929) considered C. euleri to be a species of southern South America, and julieni a subspecies of C. americanus in northern South America. Banks (1988b) noted that euleri and julieni were the same kind of bird, and recognized them as a species distinct from C. americanus. The holotype of C. euleri is ZMB 21007.

Measurements and weights Wing, M (n ⫽ 24) 119–135 (128.2 ⫾ 4.3), F (n ⫽ 9) 130–138 (133.0 ⫾ 3.0); tail, M 116–140 (125.7 ⫾ 7.5), F 131–138 (133.4 ⫾ 2.4); bill, M 22.1–26.4 (24.3 ⫾ 1.2), F 23.5–29.3 (25.9 ⫾ 1.8); tarsus, M 18.2–25.4 (21.5 ⫾ 1.7), F 23.4–25.7 (24.4 ⫾ 1.8) (AMNH, ANSP, CM, FMNH, MCZ, POC, UMMZ, USNM, ZMUC). Weight, M (n ⫽ 2) 45–53.5 (49.3), F (n ⫽ 2) 54–61.0 (57.5), F laying (n ⫽ 1) 61.4, U (n ⫽ 1) 45.0 (Novaes and Cunha Lima 1998, Magalhães 1999, ANSP, FMNH, RMNH). Wing formula, P8 ⬎ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10. Birds breeding in northern South America may average smaller than birds in southern South America.

Field characters Overall length 28 cm. Slender brown cuckoo, brown above and white below, long tail with large white tips and white edge, bill yellow at base with dark tip, wings brown with inner vane of primaries white (wing rufous in Yellow-billed Cuckoo C. americanus, wing brown in Mangrove Cuckoo C. minor).

Voice Slow series of “tówlp” notes, given 1 per sec, the note lasting 0.2–0.3 sec, beginning at 1 kHz and falling to 0.5 kHz, with the first part of the note held nearly on one pitch then dropping halfway through the call and held on a lower pitch as it fades away, in a two-part sound. Another call is a short accented

346 Pearly-breasted Cuckoo Coccyzus euleri rattle “KIkikiki” of 4 or 5 notes, the first note louder and higher in pitch (1.8 kHz at peak amplitude) and the later notes softer and dropping in pitch through the series (Hilty and Brown 1986, Hardy et al. 1990).

Range and status South America east of the Andes. Breed in northern South America and in southern Brazil and Paraguay. Occur from northern South America in Colombia (Dpto. Bolívar: Cartagena), Venezuela (Distrito Federal, Territoro Amazonas, Bolívar (Altagracia, Caicara, Upata), Apure) the Guianas (Guyana: Demerara River; Suriname: Paramaribo, Zanderi) (Penard and Penard 1910, Pinto 1938, Phelps and Phelps 1958, Hilty and Brown 1986, Haverschmidt and Mees 1994, Tostain et al. 1992), and Brazil (Amazon basin in Roraima (Igarapé Serrinhia, Colonía do Apiaú: October, December), Amazonas (Tonantins, Rio Solimoes; Rosarinho: April, June, August), Rondônia (Rio Madeira), Maranhão (Rosario: May), Ceará ( Juaeiro do Norte: December)), Pará (Santarem, Rio Tapajos, Belem, Amapa, Arucaua: April, May, July, August), and Bahia (Rio Gongogi) (Pinto 1938, Pinto and Camargo 1956, Straube and Bornschein 1989, 1995, Novaes and Cunha Lima 1998, Borges et al. 2001;AMNH, CM, FMNH, MCZ, USNM), Mato Grosso, Mato Grosso do Sul (Cantagallo), Paraná, Minas Gerais, Rio de Janeiro and São Paolo (Dubs 1992, Sick 1993, Parker and Goerck 1997). In tropical northern South America, the cuckoos are considered resident. Most records are from April to September. The birds are rare in Colombia, local in Venezuela and French Guinea, and rare in

?

Suriname (Caribbean coast near Cartagena and extreme SE Guainía along Río Negro) (Meyer de Schauensee and Phelps 1978, Hilty and Brown 1986, Haverschmidt and Mees 1994, Stotz et al. 1996). In Brazil their distribution and status are not well known. In northern Brazil there are records for most of the year, and breeding is known in October (Sick 1993, FMNH). In the Amazon basin, Roraima and Belém, the cuckoos occur from April to September and in October and December; in Cereá and Bahia they occur in May, June, October and December (Pinto 1938, Cohn-Haft et al. 1997, Novaes and Cunha Lima 1998). In eastern and southern Brazil they are uncommon breeding birds in summer (Sick 1993, Magalhães 1999); in São Paulo state their numbers are much reduced from a few decades earlier (Willis and Oniki 1992). In temperate southern South America from southern Paraguay to northern Argentina (Misiones), the birds are austral summer breeding visitors; records south of 20°S are from October to January (Hellmayr 1929, de la Peña 1986, Capper et al. 2001). In Paraguay they occur in tall humid subtropical forest near Jejui-mi in Dpto. Canindeyú in October and January (Capper et al. 2001) and one was taken near the Mennonite Colony in Dpto. Presidente Hayes c. 265 km west of Puerto Casado on 14 Oct 1936 (UMMZ 92993). In other parts of South America their status is not well known, and the cuckoos may occur mainly in migration. In Ecuador they are reported as rare migrants (specimen in Prov. Esmeraldas in September, observations in March and April); (Ridgely and Greenfield 2001, ANSP). In Bolivia they occur in Tarija in April and Andres Ibañez in May (Laubmann 1939, Remsen et al. 1986, Remsen and Traylor 1989).Accidental, a record (20 Oct 1863, AMNH 44495 holotype of “julieni”) on Sombrero I,West Indies (Lawrence 1864, Greenway 1978). Pearly-breasted Cuckoos have been confused with Yellow-billed Cuckoos in field observations, and they are not well known.

Habitat and general habits Tropical lowland evergreen forest, gallery forest, secondary forest, open woodland, sandy woodland and scrub, fruiting trees; sea level to 400 m. Occurs

Mangrove Cuckoo Coccyzus minor 347 in the canopy, alone and in mixed-species feeding flocks (Cohn-Haft et al. 1997). In southeastern Brazil, eastern Paraguay and northeastern Argentina the cuckoos live in the Atlantic forest, one of the most endangered habitats in the Neotropics and one of the most threatened in the world (Stotz et al. 1996, Madroño et al. 1997).

Food Insects, mainly caterpillars (Novaes and Cunha Lima 1998).

Breeding In Venezuela adult males taken in Distrito Federal, Territorio Amazonas and Terr. Bolívar from April

to June had large testes, a female taken at Caicara, Terr. Bolívar, in June was in breeding condition (Cherrie 1916), a male taken in Terr. Bolívar in June sang vigorously and continually, and a grown juvenile, and a nearly-grown juvenile was taken in Amazonas in August (POC). An incompletely grown juvenile was taken in Cartagena, Colombia, in January (CM), in Brazil breeds in Roraima in October (FMNH) and in the Orinoco basin in June (Cherrie 1916, Snethlage 1928). In Paraguay one bird approached another holding a small twig in the bill, then they copulated, in late October (Capper et al. 2001). Nest and eggs are undescribed, clutch 2, indicated by a specimen label that describes 2 ovulated follicles (FMNH 323751).

Mangrove Cuckoo Coccyzus minor (Gmelin, 1788) Cuculus minor Gmelin, 1788, Systema Naturae, 1(1), p. 411. [Cayenne ⫽ French Guinea] Other common name: Maynard’s Cuckoo Monotypic.

Description ADULT: Sexes alike, crown gray, back and rump grayish brown, wing brown, tail long, graduated, dark with large white oval tips, below with large white tips (T1 olive brown and without white tips, T2 olive brown with inconspicuous white tip, truncate white pattern across feather T3 to T5), edge of tail black; face with broad black line behind the eye; underparts, throat to breast whitish buff to rufous buff, belly to under tail coverts buff to rusty brown, under wing coverts whitish buff to rufous buff; eye-ring gray to yellow, iris dark brown, bill black above, yellow or orange below, feet gray. JUVENILE: Wing coverts and flight feathers brown edged buff to rufous, rectrices more pointed and tail spots less distinct than in adult, dark line through the eye indistinct, breast buffy white, bill dark above and pale below. NESTLING: Undescribed.

SOURCES: AMNH, BMNH, CM, FMNH, LSU, MCZ, MVZ, ROM, UMMZ, USNM.

Subspecies Several subspecies have been described, based on small samples (Ridgway 1916, van Rossem 1934). Larger series show no consistent differences in color or size among islands of the West Indies, or between islands and continental populations (Banks and Hole 1991). Birds in the Bahamas are pale and birds in Mexico and the Windward Islands are dark, although variation in color of the upperparts and the intensity of buff on the underparts is not strictly geographic.

Measurements and weights Bahamas and Hispaniola: Wing, M (n ⫽ 8) 125–147 (137.3 ⫾ 7.7), F (n ⫽ 9) 125–144 (131.8 ⫾ 6.9); tail, M 154–176 (163.3 ⫾ 7.8), F 157–169 (163.1 ⫾ 4.6); bill, M 25.8–31.1 (29.0 ⫾ 1.6), F 27.0–31.7 (28.0 ⫾ 4.3); tarsus, M 23.1–30.5 (27.1 ⫾ 2.4), F 25.0–29.0 (27.4 ⫾ 1.0) (FMNH); all regions: Wing, M (n ⫽ 158) 126–156 (136.5 ⫾ 5.6), F (n ⫽ 149) 124–161 (136.9 ⫾ 5.7) (Hughes 1997b). Weight, M (n ⫽ 9) 51.5–76.6 (64.3), F (n ⫽ 11) 62.5–75.0 (67.0) (Hughes 1997b). Wing formula, P7 ⬎ 8 ⬎ 6 ⬎ 9 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

348 Mangrove Cuckoo Coccyzus minor

Field characters Overall length 34 cm. Cuckoo with brown upperparts, brown wing, a broad black streak through the eye, belly and under tail coverts rufous, bill bicolored black and yellow. Adult differs from adult Yellow-billed Cuckoo C. americanus by the brown (not rufous) wing, black mask, the darker, rufous underparts, and the dark (not white) edge of outer tail feathers. Juvenile differs from young Yellowbilled Cuckoo C. americanus in the pale edges of the wing coverts and the dark edge of the tail, and from young Pearl-breasted Cuckoo C. euleri by larger size and dark edge of the tail.

Voice Low, guttural or nasal, resonant “gawk gawk gawk gawk gaawk gaawk” at 1 kHz with an overtone at 2 kHz (4 notes in 1 sec), a low (0.8 kHz) “coo-coocoo-coo” (5 notes in 2 sec), and a “whit!”. A “chakook” and a “cha-gook-chook” alarm are given near the nest, and other calls include a “cluck” and squirrel-like notes (Bendire 1895, Bent 1940, Peterson 1980, Stiles and Skutch 1989, Hardy et al. 1990, ffrench 1991, Howell and Webb 1995, Hughes 1997b, Raffaele et al. 1998, Reynard and Sutton 2000). Calls are lower in pitch than of Yellow-billed Cuckoo.

Range and status Mainly Gulf of Mexico and the Caribbean, local in western Mexico. In Florida, southern Mexico from Sinaloa and Tamaulipas to Yucatan and Oaxaca, Central America, western Panamá, northern South America from Colombia, Venezuela, the Guianas, Trinidad (Caroni swamp), Amazon estuary in NE Brazil (Belcher and Smooker 1936, Pinto 1938, Griscom and Greenway 1941, Paynter 1955, Phelps and Phelps 1958, Voous 1983, Hilty and Brown 1986, Binford 1989, ffrench 1991, Tostain et al. 1992, Rappole et al. 1993, Sick 1993, Haverschmidt and Mees 1994, Hughes 1997b, AOU 1998, Raffaele et al. 1998). They occur throughout the West Indies, the Bahamas and the Lesser Antilles from Guadeloupe to Grenada (not in Barbados) and are common in Jamaica and Puerto Rico, uncommon in Cuba in eastern coastal areas and

cays, common in Caribbean Mexico and Central America to Pacific western Panamá and the Guianas to Trinidad, and southern Caribbean islands from Netherlands Antilles to Monos, and Venezuela and Colombia (Benito-Espinal 1990, Wallace et al. 1996, Stotz et al. 1996, Hughes 1997b, Raffaele et al. 1998).They are year-round residents in most of their range (Gosse 1847, Thurber et al. 1987, Binford 1989, Raffaele et al. 1998, AOU 1998). In the southern tip of Florida and the Florida keys a few birds winter in mangroves and inland hardwood hammocks, while others arrive in spring in late March and depart in late September (Stevenson and Anderson 1994, Hughes 1997b). Odd birds occur north to coastal Texas and peninsular Florida and south to central Panamá (canal area and Pearl Islands). In western Mexico, they are common in the hills between Alamos and the Sonora-Sinaloa border (Russell and Monson 1998). Uncommon on the Florida coast and in the keys, where the species is in harm’s way from human development in its mangrove habitat; the birds persist where this habitat is protected (Hughes 1999b). Cuckoos on islands may disappear and then recolonize from other island sources. The numbers on St. Croix declined by more than 20% the year after a hurricane passed through in 1989 (Wauer 1996).

Habitat and general habits Avicennia mangroves, estuaries and thickets near water. In Florida and Mexico they occur mainly in mangroves, but in western Mexico and the Greater

Mangrove Cuckoo Coccyzus minor 349 Antilles where they occur together with Yellowbilled Cuckoo C. americanus they are not restricted to mangroves; both cuckoo species occur in scrubby riparian woodland, dry scrub and lightly wooded tropical deciduous forest, and dry hillside forest (Lack 1976, Hilty and Brown 1986). They live in dense scrub and woodland in the Bahamas, in mangroves, xeric coastal scrub, thatch palm and guinea plum in Cuba where mainly on SE coast and more common on keys on N and S coasts, in south coast swamps and gardens and dry limestone forest and lowland secondary forest in Jamaica (Lack 1976, Downer and Sutton 1990, Hughes 1997b), from semi-arid thorn scrub to dense lowland jungles in Hispaniola (Wetmore and Swales 1931), in mangroves, coffee plantations, forests and thickets but not in high mountains in Puerto Rico (Wetmore 1927), in understory and montane rainforest in the Lesser Antilles (Terborgh et al. 1978) and at 1100 m in cloud forest in El Salvador (Dickey and Van Rossem 1938).The cuckoos occur in mangroves on the Gulf coast of Mexico, as well as along river valleys into Prosopis mesquite and cactus scrub more than 100 km from the coast, and in interior tropical thorn scrub in Sonora (Eaton and Edwards 1947, Russell and Monson 1998). They live in mangroves, swamp forest and tropical deciduous forest on the Pacific coast of Mexico (Selander and Giller 1959, Haverschmidt and Mees 1994, Howell and Webb 1995), and in mangroves in Honduras and Suriname. Lowlands near sea level, the cuckoos occur locally to 1100 m, and breed in semi-arid habitat far from water where grasshoppers are abundant, to as high as 1300 m (Rowley 1984).They perch and scan the vegetation, motionless, skulking and secretive, then pursue the prey with hops and runs, and sometimes feed on the ground (McNair 1991).

Food Insects, mainly caterpillars and grasshoppers (Acrididae), other large insects including crickets, mantids, walking sticks, roaches, earwigs, cicadas, stink bugs, beetles, sugar cane borers; spiders, snails, lizards, perhaps eggs and nestlings of small birds, and fruit (Gosse 1847, Wetmore 1916, Danforth 1925,

Bent 1940,Voous 1957, Lack 1976,Wunderle 1981, Stiles and Skutch 1989).

Displays and breeding behavior In courtship feeding, the male brings an insect to the female, which she takes.The behavior may keep the pair together, it can occur without copulation, and copulation can occur without courtship feeding (Langridge 1990, McNair 1991). In copulation the female perches and elevates her bill, upper body and tail, then pumps her tail up and down 180° while giving soft calls for a minute or two. The male flies to the female without calling, grasps her bill, and the pair copulates for c. 6 sec (McNair 1991). Both parents feed the young (Bradley 1985, Stevenson and Anderson 1994).

Breeding In southern Florida they breed from May to July (Hughes 1997b), in the West Indies from February to June (Raffaele et al. 1998), in the Bahamas in May (Buden 1987), in Jamaica from March to August (Downer and Sutton 1990, Hughes 1997b), in Puerto Rico from March to October (Wetmore 1916), on Hispaniola in July (Wetmore and Swales 1931), in Lesser Antilles (St. Vincent, Grenadines) from March to July (USNM), in Trinidad from July to September ( ffrench 1991). In Mexico, they breed in Sonora with the late summer rains in August (Russell and Monson 1998) and in Oaxaca in June and July (Binford 1989); in El Salvador they breed in May (Thurber et al. 1987), in Panamá birds are in breeding condition in May (Wetmore 1968) and in Suriname they breed in June (Penard and Penard 1910, Hellebrekers 1942). The nest is a flat platform of sticks and leaves, 2–3 m above water in mangroves or in a fork of a tree above the ground. Eggs are pale greenish to bluish green, fading to greenish yellow, to nearly white in museum collections, smooth and without gloss, 31 ⫻ 23 mm, weight 0.49 g, clutch 2–3(4) (Bendire 1895, Belcher and Smooker 1936, Bent 1940, Hellebrekers 1942, Stockton de Dod 1981, ffrench 1991, Stevenson and Anderson 1994, Hughes 1997b). Incubation and fledging periods are unknown.

350 Cocos Cuckoo Coccyzus ferrugineus

Cocos Cuckoo Coccyzus ferrugineus Gould, 1843 Coccyzus ferrugineus Gould, 1843, Proceedings of the Zoological Society of London, 1843, p. 105. (Cocos Island) Monotypic.

from Dark-billed Cuckoo C. melacoryphus in rufous not dark brown wings, rufous plumage above and below, underside of tail with white tips in a continuous white pattern, and a bicolored bill.

Description

Voice

ADULT: Sexes alike, forehead and crown gray, back rufous brown, wing rufous, primaries with inner edge rufous and tips gray-brown (primaries brown in C. minor), tail brown with outer feathers nearly black (T1 grayish brown, with 1 mm whitish edge; T2 to T5 blackish brown, with broad white spots seen from below, T5 with white outer edge and inner vane), face with broad line behind the eye; underparts rufous buff to tan, under tail coverts darker rufous buff, under wing coverts rufous buff; eye-ring yellow, iris dark brown, bill black above, yellow below with black tip, feet dark gray.

A deep, dry cough “kcha” repeated 5–8 times, also a resonant guttural “k’k’k’k’k’ru’hoo” (Slud 1967, Stiles and Skutch 1989).

JUVENILE: Similar to adult, with tail pattern indistinct, T5 with whitish outer edge and inner vane with white edge, tail feathers narrower than in adult. NESTLING: Undescribed. SOURCES: AMNH, BMNH, CM, FMNH, MCZ, USNM.

Measurements Wing, M (n ⫽ 13) 126–137 (131.5 ⫾ 3.6), F (n ⫽ 7) 131–137 (134.2 ⫾ 1.7); tail, M 152–177 (163.1 ⫾ 5.8), F 158–164 (161.2 ⫾ 3.0); bill, M 30–32.3 (31.0 ⫾ 0.7), F 26.4–31 (28.5 ⫾ 2.0); tarsus, M 22.1–28.4 (26.6 ⫾ 2.3), F 27.5–28.6 (28.1) (AMNH, BMNH, CM, FMNH, MCZ, USNM). Wing formula, P8 ⱖ 7 ⬎ 6 ⬎ 9 ⱖ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 32 cm.The only cuckoo on Cocos I, it looks like a large Mangrove Cuckoo C. minor with a more rufous back and rufous wings. If differs

Range and status Cocos I (Costa Rica). Resident.This cuckoo is the least common native landbird on Cocos I, which has an area of 47 km2 with four endemic species of birds found nowhere else.Threats to the long-term existence of the cuckoo include rats, and cats, and overgrazing and habitat destruction by the introduction of deer, pigs and goats (Collar et al. 1994, Stotz et al. 1996).

Habitat and general habits Tropical lowland evergreen forest, widespread in Hibiscus thickets, tangles, hanging vines along streams, and second growth forest.The island is well forested; cuckoos forage in dense vegetation, where they hop and run, then flap and glide in flight. The cuckoos are the least tame native landbird on the island (Slud 1967, Stiles and Skutch 1989, Stotz et al. 1996).

Black-billed Cuckoo Coccyzus erythropthalmus 351

Food

Breeding

Large insects (sphingid caterpillars, cicadas), Anolis lizards (Stiles and Skutch 1989).

Not described.

Black-billed Cuckoo Coccyzus erythropthalmus (Wilson, 1811) Cuculus erythropthalma Wilson, 1811, American Ornithology, 4, p. 16, pl. 28, fig. 2. (no locality given, probably near Philadelphia, Pennsylvania). Monotypic.

Description ADULT: Sexes alike, upperparts olive brown, wing primaries brown, tail long, graduated, gray with black subterminal bands and narrow white tips below (T1 olive brown without band or tip, T2 olive brown with inconspicuous white tip,T3 to T5 with white tips 4–6 mm in length, less than the width of the rectrices), T5 brown (not white) on outer vane, underparts white from throat to under tail coverts, under wing coverts buffy white; eyering red in breeding grounds and yellow in winter, iris dark brown, bill black, feet dark gray. JUVENILE: Above olive brown, feathers narrowly edged buff to whitish on the crown, wing coverts, back and rump, wing brown to rufous with a dark leading edge, tail with rectrices narrow and tapered near tip, T1 without marks, T2 to T5 tip 2–3 mm slightly paler than base of rectrices; underparts whitish with buff to gray wash on throat; eye-ring gray changing to buff, then yellow or pale green in autumn, iris dark brown, bill black above and gray below, turning gray from the base. Buff-edged juvenile wing coverts are sometimes retained into the first breeding season. NESTLING: Skin blackish, nearly naked at hatching, with white hair-like down, the frontal apterium as dark as the bill and nearly the color of the surrounding skin; corner of gape unswollen and red, tomia white, mouth lining red, palate red with five white papillate patches (the lateral patches larger than the medial patches; the anteromedial patch

sometimes divided by the narrow rostral slit of the choana), two pairs of large elongate mediolateral patches along the pars caudalis of the choana, tongue red with a pair of small white papillated patches on the alae linguae, a large white patch of papillae on the dorsal surface and a black tip, feet blue-gray (Roberts 1936, Nolan 1975, UMMZ). SOURCES: AMNH, BMNH, CM, FMNH, LSU, MCZ, MVZ, ROM, UMMZ, USNM.

Measurements and weights Michigan: Wing, M (n ⫽ 12) 133–142 (135.2 ⫾ 2.5), F (n ⫽ 12) 133–145 (139.2 ⫾ 4.1); tail, M 142–161 (150.9 ⫾ 5.3), F 144–160 (154.4 ⫾ 5.0); bill, M 21.1–22.6 (22.0 ⫾ 0.4), F 20.7–25.3 (22.6 ⫾ 1.2); tarsus, M 20.2–24.8 (22.9 ⫾ 1.6), F 21.6–24.9 (23.4 ⫾ 1.1) (UMMZ); Weight, Michigan, breeding season: M (n ⫽ 21) 41.6–50 (46.6), one 34.0 ( found dead in June), F (n ⫽ 19) 46.2–62.8 (54.1) (birds with brood patch from 49.0, birds with egg in oviduct from 53.6 to 63.0) (UMMZ). Wing formula, P8 ⫽ 7 ⬎ 6 ⬎ 9 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 30 cm. Slender cuckoo, brown above and white below with long graduated tail, gray not blackish below, adult with no rufous in wing and small white spots on tips of outer tail feathers (underside of tail black with large white spots in Yellow-billed Cuckoo C. americanus), edge of tail dark (white in Yellow-billed Cuckoo), bill black, narrower than Yellow-billed Cuckoo. Juvenile sometimes has rufous in wing, and it has a pale buff edge of upper wing coverts. Nestling has white natal down (gray in Yellow-billed Cuckoo).

352 Black-billed Cuckoo Coccyzus erythropthalmus

Voice A rhythmic “cu cu cu, cu cu cu, cu cu cu”, and a longer and slower “kow kow kow”. Each note “cu” begins at about 1.2 kHz and drops to 1 kHz in pitch, the notes about 0.1 sec long with a slightly shorter interval between notes in a phrase, and the phrases about 1 sec apart. Black-billed Cuckoos tend to group their “cu” or “kow” calls in 2’s, 3’s and 4’s, differing from Yellow-billed Cuckoo calls which are less distinctly phrased and are guttural and lower in pitch (Hardy et al. 1990). Other calls include a croak (Hughes 2001). Nestlings beg with a buzz. Silent on the wintering grounds.

Range and status North America from southern Canada and the United States east of eastern Rocky Mountains, in the west to Alberta and Manitoba, Montana and Colorado, in the Great Lakes region and in the northeast from Ontario, southern Quebec and Nova Scotia through New England and south through the Appalachian Mts to North Carolina and Tennessee. In the southern Great Plains they breed southward to northern Oklahoma; they have bred in southern Texas (once) and they breed in the Atchafalaya River basin in southern Louisiana. Occasionally they appear and breed locally in northwestern North America (Sutton 1967, Fischer 1979, Muth 1991, Nicholson 1997, AOU 1998, Hughes 2001, Graves et al. 2002). Migratory, they move at night. Spring migration records in Florida are in April and May, autumn migration from late August to late October. Spring migrants occur from early April to as late as mid-June in Louisiana (LSU). In Mexico an uncommon or rare migrant mainly from Guerrero, Guanajuato and Tamaulipas southward, mostly in lowlands of Gulf and Caribbean (AOU 1998) mainly in September and April and May (six localities in Oaxaca, Binford 1989), in Central America and Panamá from late September to early November and April to early May (Willis and Eisenmann 1979, Stiles and Skutch 1989), on Cuba and Hispaniola a regular but rare migrant in autumn, less regular on the Bahamas and other islands of the West Indies (Raffaele et al. 1998,

Bradley 2000). A few occur west to the Pacific from SE British Colombia south to Sinaloa.An old record from Trinidad was in September (Léotaud 1866), although birds have not appeared there in recent years ( ffrench 1991). Migrant in passage in northern South America, birds pass further south to winter (Rodner et al. 2000). In western Venezuela (Táchira, Aragua) and Colombia (both east and west of the Andes) birds are mainly in passage and wintering birds are not known for certain (Meyer de Schauensee 1978, Hilty and Brown 1986, POC, USNM). In Ecuador, these cuckoos occur from early November to April with most records late in this period, perhaps in spring migration from further south (Ridgely and Greenfield 2001). Winter range is in South America. In Ecuador cuckoos occur in winter east of the Andes at Podocarpus NP, on the Pacific slope at La Caronia, El Oro, and in the SW Ecuadorian lowlands, and in western Peru they occur south to Trujillo (Fjeldså and Krabbe 1990, Ridgely and Greenfield 2001, J. Fjeldså, in litt., ZMUC). East of the Andes they extend much further south. Most winter records (December to February) are in eastern Peru

Black-billed Cuckoo Coccyzus erythropthalmus 353 of the species range (Peterjohn et al. 1995). Sneaky birds, usually stay in dense cover.

?

Habitat and general habits

(Amazonas, Puno, Lambayeque, Pasco, San Martín, Ucayali and Madre de Dios) and Bolivia (Cochabamba and Santa Cruz) (Ridgway 1916, Remsen and Ridgely 1980, Remsen and Traylor 1983, 1989, Fjeldså and Krabbe 1990, Rappole et al. 1993, Ridgely and Greenfield 2001, FMNH, LSU). In Brazil, the species is hardly known; it occurs in the extreme west (Rio Juruá) (Whittaker and Oren 1999). Rare in Paraguay (Alto Chaco) (Short 1976, Hayes 1995), and there is a winter record in northern Argentina (Misiones) (Meyer de Schauensee 1970, de la Peña and Rumboll 1998). Winter records are in lowlands. In Ecuador occasional birds perhaps in passage are seen at 2800 m near Quito and 3950 m at Papallacta Pass, and there is one record in the Galapagos (Harris 1982). Black-billed Cuckoos are vagrants in autumn to western Europe, where they are accidental in the Azores, Iceland, Greenland, Britain and Italy and are less common than Yellow-billed Cuckoos (Cramp 1985, ZMUC). Local breeding densities vary from year to year, the birds appearing in numbers in some years with outbreaks of caterpillars. Cuckoos increased somewhat in the period 1965–1979, when the peak years coincided with widespread outbreaks of caterpillars (Robbins et al. 1986, Eastman 1991). Continent-wide population numbers showed little change in the period 1966–1993 (Price et al. 1995) when there were regional trends of ⫹ 0.9% per year in the east, ⫹ 0.5% per year in central North America, and ⫺ 1.1% per year in the western part

Forest, both deciduous and coniferous, open woodland, dense deciduous thickets, willow, alder, aspen and vines, hawthorn meadows, shrubby fields and pastures. In breeding season they occur together with Yellow-billed Cuckoos through much of the range, but are more northern; they are in more extensive forests and in the Appalachians they are at higher elevations. They are the more numerous cuckoo in the northern part of their common range. In migration and in winter they are in tropical evergreen forest to arid subtropical scrub, shrubs and mangroves; from lowlands to 2000 m. They forage in tree foliage, skulk about quietly, cock the head, wait until an insect moves, then grab it. They also fly to branches and tree trunks and snatch insects from the surface. They wipe a caterpillar on a branch, shake it to remove the guts, then toss the insect into the air and catch it with open bill and swallow it. They feed mainly in trees and also on the ground (Forbush 1927, Bent 1940,Agro 1994, Hughes 2001). In winter, they join mixedspecies feeding flocks (Terborgh et al. 1984, Robinson et al. 1995).

Food Insects, mainly caterpillars, including deforesting pests (gypsy moths Lymantria dispar, fall webworms Hyphantria cunea, army worms Pseudaletia unipunctata, tent caterpillars Malacosoma spp.); larvae of sawflies, grasshoppers, crickets, bugs and beetles, periodic and annual cicadas Magicicada spp. and Tibicen spp., also spiders and phalangids; occasionally small tree frogs, small fishes, and eggs of other birds.The cuckoos feed on berries and other fruits in autumn (Beal 1898, Forbush 1927, Squires 1930, Bent 1940, Witter and Kulman 1972, Hughes 2001).

Displays and breeding behavior Monogamous. The birds occur in pairs and are territorial in the breeding season. In courtship

354 Gray-capped Cuckoo Coccyzus lansbergi feeding, the male calls with a caterpillar in his bill, the female spreads and flicks her tail and makes a mewing sound, then the male gives her the caterpillar. Copulation occurs near the nest and sometimes continues through the incubation period (Spencer 1943).

Breeding In eastern North America the birds breed from May to September (Bendire 1895), long enough to rear two broods in some seasons. In Oklahoma they breed in May and June (Sutton 1967), in Minnesota and Kansas from May to August (Roberts 1936, Johnston 1964), in Maryland from May to July (Stewart and Robbins 1958), and in Michigan and Ontario from late May to early September (Wood 1951, Peck and James 1983). The nest is a shallow cup of sticks, twigs, inner bark, fine rootlets and weed stems, lined with dead and green leaves or other plant matter, built on the ground or more often on horizontal limbs in evergreens or deciduous trees and shrubs, to 2 m or higher above ground (Bendire 1895, Spencer 1943, Johnston 1964, Peck and James 1983, Gauthier and Aubry 1996). It is 12.5 to 15 cm diameter, with outside depth 5 to 6.5 cm. Both sexes build the nest. Eggs are light blue (darker than Yellow-billed Cuckoo), smooth, not glossy, 27 ⫻ 21 mm. The clutch is 2–5, occasionally 6–8 when more than one female lays in a nest; in Kansas 2–3 (2.5). The number of eggs is larger in seasons with a big emergence of insects such as tent caterpillars, either because the female responds to the extra food, or because more than one female lays in a nest. The eggs are laid at intervals of 1–3 days. Both parents incubate; the average time a bird is on the nest is 90 minutes. Incubation begins with

the first egg. Nestlings hatch asynchronously; the incubation period for an egg is 10–11 days. The parents leave the hatched shell in the nest where small bits remain through the period of parental care. The brood is fed by both parents, who carry the nestlings’ fecal sacs from the nest or swallow them. Nestlings beg in a vertical posture, with neck stretched upwards and wings stretched straight out to the side and flapped (Figure 4.12). Nestlings can grasp the nest lining a few hours after they hatch, and they give a bee-like buzz begging for food. They are 8.5 g on the day after they hatch, grow to 26 g at 6 days, and fledge at 8–9 days, or even earlier if disturbed. Nestling is covered with feather sheaths by day 6, these burst and the bird is feathered on day 7. Because they hatch on different days and develop rapidly, in a large clutch or brood the first nestling can fledge before hatching of the last egg.Weights of four well feathered, stub-tailed birds out of the nest were 17.9, 21.5, 22.3 and 33.3 g (UMMZ), average 23.8 g. Nest attendance, at a brood in the nest of a color-marked pair in Michigan the pair switched attendance at 2-hour intervals through the day (Bendire 1895, Herrick 1910, Roberts 1936, Bent 1940, Spencer 1943, Johnston 1964, Oberholser 1974, Sealy 1978, 1985, Peck and James 1983, Hughes 2001, RBP). Occasionally Black-billed Cuckoos lay in the nests of other Black-billed Cuckoos, Yellow-billed Cuckoos and other kinds of birds (Bendire 1895, Bent 1940, Spencer 1943, Nolan and Thompson 1975, Peck and James 1983,Thomas 1995, Hughes 1997a, 2001). Breeding success, in Michigan, of 18 eggs laid in six nests, 14 hatched and 10 young fledged from the nest, a survival of 55% from egg to fledging (Spencer 1943).

Gray-capped Cuckoo Coccyzus lansbergi Bonaparte, 1850 Coccyzus lansbergi Bonaparte, 1850, “Forsten” Conspectus Systematis Ornithologiae, 1, p. 112. (Santa Fé de Bogotá [Colombia]) Monotypic.

Description ADULT: Sexes alike, crown and nape dark gray, back and rump rufous brown, wing brown, tail long and black, the central tail feathers T1 dark to tip, T2 to

Gray-capped Cuckoo Coccyzus lansbergi 355 T5 with large rounded white tips, face dark gray and cheeks darker gray; underparts rufous from chin to belly and under tail coverts, the breast darkest, under wing coverts rufous; eye-ring gray to yellowish or white, iris brown, bill black with yellow or gray spot at base of lower mandible, feet blackish. JUVENILE: Similar to adult, but crown brown and pale tips of rectrices less distinct than in adult. NESTLING: Hatchling naked, skin black or blackish (Marchant 1960). SOURCES: AMNH, ANSP, BMNH, FMNH, LSU, MCZ, MVZ, USNM, ZMA, ZMUC.

Measurements and weights Wing, M (n ⫽ 8) 107–118 (113.9 ⫾ 4.2), F (n ⫽ 8) 110–117 (113.2 ⫾ 1.6); tail, M 117–156 (130.0 ⫾ 14.0), F 120–142 (137.2 ⫾ 4.6); bill, M 22.6–27 (24.1 ⫾ 1.7), F 21.4–27.0 (23.8 ⫾ 2.4); tarsus, M 22.4–28.5 (25.9 ⫾ 1.6), F 24.0–26.0 (24.4 ⫾ 0.8) (AMNH, ANSP, CM, LSU, MCZ). Weight, M (n ⫽ 4) 46–56 (50.3); F (n ⫽ 1, emaciated, Bonaire), 27.9 (ANSP, LSU, MCZ, ZMA). Wing formula, P7 ⬎ 8 ⬎ 6 ⬎ 5 ⬎ 9 ⱖ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 26 cm. A small dark cuckoo with a gray cap, rufous upperparts and dark rufous underparts.

Voice Fast series of 6–8 mellow “cu” notes,“cucucucucucucu-cu” with a pause before the last note, each note at 0.8 kHz and falling near the end, the note 0.1 sec long and the notes repeated 7 per sec; and a low groan when disturbed (Hilty and Brown 1986, Hardy et al. 1990).

Range and status South America in northern Colombia (La Guarija (Potrero de Venancio), Magdalena (Mamatoco, Magdalena Valley, Santa Marta, Rio Frío, Tucurinca), Atlántico, Bolívar (Cartagena, Simití,

Turbaco) and Norte de Santander (Villa Felisa)), northern and western Venezuela (Zuila west of Lake Maracaibo, Mérida, Larea, Carabobo, Aragua, Distrito Federal, Miranda), western Ecuador (Manabí, Los Riós, Guayas (Daule, Guayaqul, Punta Sta Elenae), El Oro (Portovelo), Loja (El Empalame) and northwestern Peru (Dpto. Lambayeque: Las Pampas near Olmas, in August: LSU 113780) (Phelps and Phelps 1958, Hilty and Brown 1986, Ridgely and Greenfield 2001). Resident, perhaps partly migratory. Observations on the east slope of the Andes at Bazea in W Napo, Ecuador, in February, and in Bosque Ampay, Apurímac, Peru, in August, suggest an Andean passage (Fjeldså and Krabbe 1990, Ridgely and Greenfield 2001) and an October record in Bonaire indicates dispersal or seasonal movement (ZMA 35.233, Voous 1983). Rare to fairly common.

Habitat and general habits Tropical deciduous forest, arid forest, mesquite thicket with dense weedy undergrowth, bushy fields, gallery forest, in mid-story and canopy; low thickets and dense shrubs near water; sea level to 800 m, in Colombia to 600 m, in Ecuador mostly below 900 m although seen to 1300 m and 1800 m, in Venezuela occasionally to 1400 m (Hilty and Brown 1986, Ridgely and Greenfield 2001, LSU). They feed in thickets and other low growth; concealed and hard to see.

Food Insects, mainly caterpillars, also wasps (Wetmore 1968, LSU).

356 Chestnut-bellied Cuckoo Coccyzus pluvialis

Breeding Season is May in Colombia, during rainy period in first half of the year in Ecuador (Marchant 1960, Hilty and Brown 1986, Ridgely and Greenfield 2001). The nest is a flat platform of sticks, lined with lichens, built 1–2 m above ground in horizon-

tal branches in a thicket. Eggs are greenish white with a rough chalky surface, 26 ⫻ 20 mm, clutch 2–3 (Marchant 1960, Schönwetter 1964). Incubation period is unknown, nestling period is within the range 8–13 days; the young leave the nest before they can fly (Marchant 1960).

Chestnut-bellied Cuckoo Coccyzus pluvialis (Gmelin, 1788) Cuculus pluvialis Gmelin, 1788, Systema Naturae 1(1), 411. Monotypic. Other common names: Old Man Bird, Hunter. Other names: Hyetornis pluvialis (Gmelin, 1758)

Description ADULT: Sexes alike, large cuckoo, crown dark graybrown, back brown, wing brown, tail long, graduated, black glossed purple with broad white tips, tail barred white from below; underparts, chin whitish, throat light gray, upper breast gray, belly and under tail coverts dark rufous; orbital skin blackish, iris red to brown, bill decurved, black with light bluish gray, mouth lining black, legs bluish gray to dark gray. JUVENILE: Similar to adult, tail dark brown not glossed and with broad white tips.

Field characters Overall length 50 cm. Large cuckoo with a white beard and a deep voice, it has a dark gray cap, brown back, long black tail with broad white tips, light breast and dark rufous belly, and decurved bill (bill slender and straight in lizard-cuckoo Coccyzus vetula).

Voice Hoarse braying raven-like “quak-quak-ak-ak-akak”, slow at beginning and accelerating towards the end, in a broad-band spectrum with most sound energy around 1.4 kHz and the notes repeated about 4 in a second. Another call is a resonant low “gwaauu”, rising then falling in pitch, and a third is a repeated hoarse croak (Gosse 1847, Hardy et al. 1990, Raffaele et al. 1998, Reynard and Sutton 2000).

NESTLING: Undescribed.

Range and status

SOURCES: AMNH, BMNH, FMNH, FU, MVZ, USNM.

Jamaica. Resident, they are suspected to move to lower elevations in winter; a few in southern lowlands occur in the same site all year (Lack 1976). Uncommon.

Measurements and weights Wing, M (n ⫽ 8) 165–189 (177.4 ⫾ 8.4), F (n ⫽ 7) 189.5–200 (190.5 ⫾ 4.8); tail, M 279–309 (290.4 ⫾ 12.4), F 273–318 (289⫾16.4); bill, M 39.2–42.0 (40.2 ⫾ 1.0), F 36.4–47.0 (42.0 ⫾ 5.3); tarsus, M 39.5–47.5 (41.8 ⫾ 3.3), F 41–44 (42.0 ⫾ 1.4) (AMNH, BMNH, FU, LSU). Weight, M, juveniles (n ⫽ 2) 129–130 (129.5) (USNM), F, adult (n ⫽ 1) 189 (MVZ), U (n ⫽ 1) 163 (Faaborg 1985). Wing formula, P6 ⫽ 5 ⬎ 7 ⬎ 4 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Habitat and general habits Tropical lowland evergreen forest, second growth forest, in thickets in open areas, in woodland in wet mountain areas and brush-covered limestone hills in cockpit country; occur from the lowlands to 1500 m (Gosse 1847, Lack 1976, Faaborg 1985, Stotz et al. 1996). They forage in mid-story and canopy, run along branches and glide on extended wings from tree to tree (Downer and Sutton 1990).

Rufous-breasted Cuckoo Coccyzus rufigularis 357

Food Large insects (katydids, caterpillars, beetles), insects that burrow in sapwood of trees or that live beneath the bark; phasmids (150–170 mm long); lizards, mice, bird nestlings and eggs (Gosse 1847, Danforth 1928, Cruz 1975, Lack 1976, Raffaele et al. 1998, LSU).

Breeding Season is March to June. The nest is a shallow saucer of twigs built in a bush or tree. Eggs are white, 40 ⫻ 31 mm, clutch 2–4. Incubation and nestling periods are unknown (Gosse 1847, Raffaele et al. 1998).

Rufous-breasted Cuckoo Coccyzus rufigularis Hartlaub, 1852 Coccyzus rufigularis Hartlaub, 1852, Naumannia, 2, p. 55. (mountain forests of Spanish Santo Domingo [⫽ Dominican Republic]). Monotypic. Other common names: Bay-breasted Cuckoo. Other names: Hyetornis rufigularis (Hartlaub, 1852).

Description ADULT: Sexes alike, upperparts head to rump gray, wing primaries chestnut with dark gray tips and dark outer and inner edges, tail long and graduated, T1 gray with black third of the tip, T2 gray basal half and black terminal half, T1 and T2 with short (3–4 mm) white tips, T3 to T5 black with broad white tips, tail black barred white from below, face gray; underparts chin to breast dark chestnut, belly pale rufous and under tail coverts buff, under wing coverts light rufous, wing rufous below; eye-lids yellow to gray, iris brown, bill narrow, straight with decurved tip, blackish above and yellowish below, feet gray. JUVENILE: Similar to adult, rectrices narrower, T1 dark gray with a rufous cast. NESTLING: Undescribed.

SOURCES: AMNH, BMNH, CM, FMNH, LSU, MCZ, MSNG, USNM.

Measurements and weights Wing, M (n ⫽ 15) 161–182 (171.6 ⫾ 8.8), F (n ⫽ 9) 167–188 (180.8 ⫾ 6.9); tail, M 243–274 (260.4 ⫾ 11.8), F 248–295 (272.2 ⫾ 13.6); bill, M 34–41 (38.4 ⫾ 3.0), F 46–50 (47.5 ⫾ 1.5); tarsus, M 35.5– 42 (39.6 ⫾ 2.0), F 40–45 (41.9 ⫾ 1.9) (AMNH, FMNH, LSU, MSNG). Weight, M (n ⫽ 1) 128 (USNM). Wing formula, P6 ⫽ 5 ⬎ 7 ⬎ 4 ⬎ 3 ⬎ 8 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 48 cm. Large cuckoo with gray upperparts, chestnut underparts, long black tail with broad white tips, and thick and decurved bill (bill is slender and straight in lizard-cuckoos Coccyzus longirostris).

Voice A strong “cua”, a call with harmonic overtones and simultaneous peak frequencies at 0.7, 1.4 and 2.1 kHz, the pitch rising to this peak, then the sound becoming more broad-band, like a bleating lamb. Also gives hoarse single calls with a wavering pitch and a guttural series “u-ak-u-ak-ak-ak-ak-akak-ak-ak” (Wetmore and Swales 1931, Hardy et al. 1990, Raffaele et al. 1998, LNS).

358 Jamaican Lizard-cuckoo Coccyzus vetula

Range and status Hispaniola and Gonâve I. Resident. Shy, rare to uncommon. Considered threatened and vulnerable due to a decline in numbers in the twentieth century.They occur in four protected areas in the western Dominican Republic, they are rare or gone in Haiti, and there are no recent records from Gonâve (Collar et al. 1994, Stotz et al. 1996). Much less common than lizard-cuckoo on Hispaniola.

Habitat and general habits Tropical deciduous forest, thorn forest, arid lowlands to mountain rain forest; lowlands to 900 m. Shy, they remain in dense cover, run along limbs of trees, forage in mid-story and canopy, creep around branches and foliage, and leap from branch to branch (Wetmore and Swales 1931, Schwartz and Klinikowski 1965,Wauer 1996).

Food Insects, including grasshoppers, mantids, grubs and caterpillars, pentastomids, beetles; also lizards, mice, small mammals, and bird eggs and nestlings

(Wetmore and Swales 1931, Stockton de Dod 1978, 1981, Raffaele et al. 1998).

Breeding A bird had an egg in the oviduct ready to lay in May; other records are from March to June (Wetmore and Swales 1931, Raffaele et al. 1998). Nest is a shallow saucer built of sticks, in a tree. Eggs are dirty white, 38 ⫻ 25 mm, clutch 2 (Schönwetter 1964, Raffaele et al. 1998). Incubation and nestling periods are unknown.

Jamaican Lizard-cuckoo Coccyzus vetula (Linnaeus, 1758) Cuculus Vetula Linnaeus, 1758, Systema Naturae (edn. 10), 1, 111. Monotypic. Other common names: Old Woman Bird. Other names: Saurothera vetula (Linnaeus, 1758).

Description ADULT: Sexes alike, crown to nape dark brown, backgray-brown, wing with rufous patch formed by primaries, tail long and graduated,T1 gray with black subterminal band and broad white tip, others black with broad white tips, white tips overlap with outer tail from below appearing white except at base; underparts, chin and throat whitish, breast and belly light rufous, under tail coverts buff, under wing coverts light rufous; orbital skin red, iris

brown, bill long, straight, black above and gray below, feet gray. JUVENILE: Similar to adult, plumage looser and fluffier, tail feathers narrow (T1 at 30mm from tip is 32mm wide in juvenile, 38mm wide in adult),T1 and T2 with no black band, T2 with white tip 8mm, secondaries with buff edge on tip. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, LSU, ROM, USNM.

Measurements and weights Wing, M (n ⫽ 10) 120–128 (123.8 ⫾ 3.0), F (n ⫽ 6) 117–136 (128.8 ⫾ 3.7); tail, M 194–203 (199.0 ⫾

Jamaican Lizard-cuckoo Coccyzus vetula 359 3.8), F 190–227 (202.0 ⫾ 13.5); bill, M 43–51 (48.2 ⫾ 2.6), F 38.7–49 (43.9 ⫾ 4.5); tarsus, M 29– 32.4 (30.2 ⫾ 1.4), F 30–32 (30.9 ⫾ 0.9) (AMNH, BMNH, FMNH, LSU, USNM). Weight, U (n ⫽ 2) 86.0–104.6 (95.3) (R. L. Sutton). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 8 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 40 cm. Large cuckoo with tail more than half the total length of the bird, a long straight bill, gray upperparts, white throat, light rufous breast and belly, and tail nearly all white when viewed from below ( plumage dark and bill curved in Jamaican Chestnut-bellied Cuckoo Coccyzus pluvialis).

Voice Call, a rapid “cak-cak-cak-ka-ka-ka-k-k” sometimes becoming faster and lower in pitch near the end, given at about 15 notes per sec and a broadband spectrum with peak amplitude around 2.6 kHz with most energy below this pitch. Another call is a single braying “gaah” (Gosse 1847, Curio 1970, Hardy et al. 1990, Downer and Sutton 1990, Reynard and Sutton 2000).

Range and status Jamaica. Endemic resident, widespread, less common in Jamaica than the larger Chestnut-bellied Cuckoo C. pluvialis (Raffaele et al. 1998).

Habitat and general habits Moist or wet mid-elevation forests, wooded ravines tropical lowland evergreen forests, forages in midstory and canopy; wooded areas and semi-arid country with trees and shrubs, lianas and briars; in wet mountain areas and brush-covered limestone hills, the “cockpit country” of west-central Jamaica; lowlands to 1200 m. The birds forage by moving slowly through the lower canopy and understory, jumping from twig to twig or climbing up the stems of young trees, stopping to perch and twisting their necks to peer about, then leaping from tree to tree, and occasionally feeding on the ground. They peer into food sites and probe for prey (Gosse 1847, Cruz 1975, Lack 1976, Downer and Sutton 1990, Stotz et al. 1996, Wauer 1996, Raffaele et al. 1998).

Food Lizards (Anolis), tree frogs (Eleutherodactylus), caterpillars, locusts, mantids, cicadas, spiders, mice and nestling birds (Gosse 1847, Cruz 1975, LSU).

Displays and breeding behavior Socially monogamous, they are seen in pairs. Male spreads his tail and wings to the female (Gosse 1847). Courtship feeding accompanies copulation (Curio 1970).

Breeding Season is March to August (Raffaele et al. 1998) and courtship was seen in October (Curio 1970). Nest is a flat platform of twigs, lined with leaves, measuring 20 cm across with a cup only 2.5 cm deep, built in a tangle of branches, twigs and bromeliads in a tree. Eggs are white (Raffaele et al. 1998) to pale blue, 32.5 ⫻ 24 mm (Schönwetter 1964). Both parents attend to the nest and feed the young, which gives a hissing call and stretches its neck upwards as it begs (Levy 1984). Incubation and nestling periods are unknown.

360 Puerto Rican Lizard-cuckoo Coccyzus vieilloti

Puerto Rican Lizard-cuckoo Coccyzus vieilloti (Bonaparte, 1819) Saurothera vieilloti Bonaparte, 1819, Nouveau Dictionnaire d’Histoire Naturelle Appliquées Arts etc (nouv. éd.), 32, 348. Monotypic.

Description ADULT: Sexes similar (males are paler, females are darker above and below), upperparts rufous brown, wing rufous brown, tail long and graduated, gray, T2–5 with a broad black subterminal band and broad white tips, tail appears barred from below; underparts, throat pale gray, breast gray, belly and under tail coverts rufous; bare skin around eye red with white crescent below eye, iris brown, bill long and straight, black above and gray below, feet gray. JUVENILE: Upperparts brown with feathers edged rufous, wing coverts and flight feathers brown edged rufous; underparts, chin and throat pale gray, belly paler rufous than in adult, tail paler with black bands narrower and tips more pointed than in adult. NESTLING: Undescribed. SOURCES: AMNH, CM, FMNH, USNM.

Measurements and weights Wing, M (n ⫽ 11) 122–138 (128.5 ⫾ 4.2), F (n ⫽ 10) 125–134 (130.4 ⫾ 3.6); tail, M 224–251 (237.8 ⫾ 9.7), F 230–267 (244.9 ⫾ 11.4); bill, M 46.9–50 (47.6 ⫾ 1.8), F 43.5–50 (43.2 ⫾ 11.3); tarsus, M 28.8–39.8 (32.1 ⫾ 3.0), F 29–36 (34.2 ⫾ 2.0) (AMNH, FMNH, USNM). Weight, M (n ⫽ 3) 74.5–83.4 (80.4), F (n ⫽ 2) 87.9–96.9 (92.4), F (laying) (n ⫽ 2) 107–110.5 (108.8) (USNM); U (n ⫽ 23) 68–104 (81.5 ⫾ 10.6) (W. Arendt). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 8 ⫽ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 42 cm. Large cuckoo with tail more than half the length of the bird, a long straight bill,

rufous upperparts, gray breast and bright rufous belly and under tail coverts.

Voice Throaty, long emphatic “ka-ka-ka-ka . . .” gradually accelerating and becoming louder, the notes repeated about 10 per sec with a peak frequency at 2.4kHz, the call lasting 6 sec or more, and soft calls “caw” (Hardy et al. 1990, Raffaele et al. 1998).

Range and status Puerto Rico. Endemic resident on Puerto Rico, and formerly on Vieques I (Wetmore 1916, 1927, Wunderle 1995, AOU 1998). A record from Virgin Islands (St. Thomas: Raffaele et al. 1998) suggests the odd dispersal.

Habitat and general habits In the haystack hills of the north coast, in mountainous areas with thick forests, and in dry coastal forests near Guánica, also swamp forests; tropical deciduous forest, tropical lowland evergreen forest, shaded coffee plantations, brush-covered limestone hills. Occur in lowlands and in mountains to 800 m (Wetmore 1916, 1927, Faaborg 1985, Stotz et al. 1996, Wauer 1996, Raffaele et al. 1998). Rare to common in wooded areas, more common at lower elevations, inconspicuous.The birds perch quietly in dense vegetation. Slow and methodical in their movements, they forage in mid-story and in the canopy and occasionally come

Cuban Lizard-cuckoo Coccyzus merlini 361 to the ground, and they have been called “ground cuckoo” (Wetmore 1916, Raffaele et al. 1998).

Food Small lizards made up 79% of the diet in one sample, also large spiders, large insects including walking sticks, cicadas, caterpillars and moths (Wetmore 1916).

Breeding There are records in April, May, September, November and December, and it perhaps breeds all year.The nest is a loose platform of sticks and leaves. Eggs are white or pale blue with a thin white outer layer, 34 ⫻ 25 mm, clutch 2–3 (Wetmore 1927, Raffaele 1989, Schönwetter 1964, Raffaele et al. 1998). Incubation and nestling periods are unknown.

Cuban Lizard-cuckoo Coccyzus merlini (d’Orbigny, 1839) Saurothera merlini d’Orbigny, 1839, in La Sagra’s Historia fisica politica y natural de la Isla de Cuba,Aves, p. 115, pl. 25. Polytypic. Four subspecies. Coccyzus merlini merlini (d’Orbigny, 1839); Coccyzus merlini bahamensis (Bryant, 1864); Coccyzus merlini decolor (Bangs and Zappey, 1905); Coccyzus merlini santamariae (Garrido, 1971). Other common names Great Lizard-cuckoo. Other names Saurothera merlini d’Orbigny, 1839.

wing smaller on average than birds in Cuba; Bahama Islands (Andros, New Providence, Eleuthera); Coccyzus merlini merlini (d’Orbigny, 1839), above olive brown; Cuba; Coccyzus merlini decolor (Bangs and Zappey, 1905); above grayish brown, central tail feathers unmarked gray to tip; Isle of Pines; Coccyzus merlini santamariae (Garrido, 1971); above pale, smaller than S. m. merlini; islands of north-central Cuba (Cayo Santa Maria, Cayo Loco).

Description

Measurements and weights

ADULT: Sexes alike, crown and back brown, wing rufous, tail long and graduated, central tail feathers T1 brownish gray to tip,T2 to T5 dark gray with a black subterminal band and broad white tip, tail appears barred from below, face with light gray cheeks continuous with gray breast; underparts, throat whitish, breast light gray, belly and under tail coverts rufous; bare skin around eye red to orange; iris gray to brown, bill long and straight, blackish above and gray below, feet gray to pink.

Cuba: Wing, M (n ⫽ 6) 163–182 (174.0 ⫾ 6.6), F (n ⫽ 6) 172–196 (180.0 ⫾ 10.3); tail, M 257–352 (312.0 ⫾ 31.5), F 280–362 (320.0 ⫾ 31.0); bill, M 49–63 (54.2 ⫾ 5.2), F 50–59.5 (54.8 ⫾ 3.4); tarsus, M 38–46.5 (41.8 ⫾ 3.0), F 40–49 (45.8 ⫾ 5.3) (AMNH). Weight, M (n ⫽ 2) 124–145 (134.5), F (n ⫽ 2) 127–186 (155) (ANSP, UMMZ, USNM). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 8 ⫽ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

JUVENILE: Tail feathers narrow, more pointed and without distinct terminal spots; iris brown.

Field characters

SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM.

Overall length 54 cm. Large cuckoo with tail more than half the length of the bird, a long straight bill, rufous wings, brown upperparts and rufous belly. Largest of the lizard-cuckoos, it glides more than it flies, and it can fly over long stretches.

Subspecies

Voice

Coccyzus merlini bahamensis (Bryant, 1864); above gray, belly buff, central tail feathers with black tip,

A throaty, long chatter “ta coo-tacoo, ka-ka ka ka ko-ko”, the notes given about 12 per sec, a peak

NESTLING: Undescribed.

362 Hispaniolan Lizard-cuckoo Coccyzus longirostris amplitude of 2.0kHz and a broad-band spectrum with most energy below 2kHz, the call lower in pitch than in other lizard-cuckoos.Another call is a loud guttural “tuc-wuh-h” or “táco . . . táco” (Gundlach 1874, Hardy et al. 1990, Garrido and Kirkconnell 2000, LNS).

Range and status Cuba, Isle of Pines, and Bahama Islands. Endemic resident. Solitary, widespread and rather common and tame in Cuba and the larger cays north of Camagüey Province; secretive and uncommon in the Bahamas (Miller 1894, Ridgway 1916, Garrido 1971,Wallace et al. 1996,Wauer 1996, Garrido and Kirkconnell 2000).

Habitat and general habits Vines and dense vegetation in tropical lowland evergreen forest, tropical deciduous forest, secondary forest, pine forest, in woodland and thickets, vines and creepers, abandoned coffee plantations, overgrown pastures and rough limestone hills; lowlands to 1200 m (Gundlach 1874, Barbour 1943, Stotz et al. 1996, Raffaele et al. 1998).They forage in the canopy, in midstory and on the ground; the flight is slow and direct.

Food Lizards, frogs, snakes, bird eggs and nestlings; insects (grasshoppers, bees and wasps) (Barbour 1943, Garrido and Kirkconnell 2000; AMNH, ANSP).

Breeding Season is April and May, rarely in October. Nest is a shallow saucer of twigs, in a tree. Eggs are white, somewhat glossy, 40 ⫻ 30 mm, clutch 2–3 (Gundlach 1874, Schönwetter 1964). Incubation and nestling periods are unknown.

Hispaniolan Lizard-cuckoo Coccyzus longirostris (Hermann, 1783) Cuculus longirostris, Hermann, 1783, Tabula affinitatum animalium olim academico specimine edita . . ., p. 186. Polytypic. Two subspecies. Coccyzus longirostris longirostris (Hermann, 1783); Coccyzus longirostris petersi (Richmond and Swales, 1924). Other names Saurothera longirostris (Hermann).

Description ADULT: Sexes similar, upperparts brownish gray, wing chestnut with rufous patch, tail long and graduated, T1 central rectrices brownish gray tipped black, other rectrices black except brown at base and with broad white tips, tail appears black with white end from below; underparts, throat

rusty (darker in female), breast pale gray, belly and under tail coverts rufous; bare skin around eye red, iris brown, bill long and straight, black above and gray below, feet gray. JUVENILE: Above gray washed brown, tail brown not gray, feathers narrower than in adults, white ends to all but central rectrices, the white tipped buff, throat buff to whitish, breast pale gray, belly buff not rufous. NESTLING: Undescribed. SOURCES: AMNH, ANSP, CM, FMNH, MSNM, MVZ, ROM, UMMZ, USNM.

Hispaniolan Lizard-cuckoo Coccyzus longirostris 363

Subspecies Coccyzus longirostris longirostris (Hermann, 1783); underparts darker, throat light rufous; Hispaniola, Saona I. Coccyzus longirostris petersi (Richmond and Swales, 1924); underparts paler, throat white; La Mahotiere, Gonâve I.

Measurements and weights Hispaniola: Wing, M (n ⫽ 11) 126–139 (134.3 ⫾ 4.4), F (n ⫽ 8) 132–142 (137.8 ⫾ 3.8); tail, M 197–241 (217.7 ⫾ 12.8), F 228–238 (232.3 ⫾ 5.5); bill, M 43–53 (49.4 ⫾ 3.4), F 40–51.5 (48.2 ⫾ 2.7); tarsus, M 31–35 (33.5 ⫾ 1.3), F 34.5–39 (36.3 ⫾ 2.1) (AMNH). Weight, Dominican Republic: M (n ⫽ 6) 83.4–99.5 (92.3), F (n ⫽ 6) 92.5–128.5 (105.4) (USNM); Saona I, U (n ⫽ 5) 84.2–97 (90.1) (W. Arendt). Wing formula, P7 ⫽ 6 ⫽ 5 ⫽ 4 ⬎ 8 ⫽ 3 ⬎ 2 ⬎ 1 ⬎ 9 ⬎ 10.

Field characters Overall length 42 cm. Large cuckoo with tail more than half the length of the bird, a long straight bill, gray upperparts, rusty throat, gray breast and rufous belly. This is the most common large cuckoo on Hispaniola; the less common Rufous-breasted Cuckoo Coccyzus rufigularis has a thick curved bill, darker throat and breast with more rufous in the wing.

Voice Throaty, rattling call, like in other lizard-cuckoos, “ka-ka-ka-ka . . . kau-kau-ko-ko”, repeated at 12 notes per sec, the notes a broad-band sound with peak amplitude at 2.4kHz and most energy at lower than 2.2kHz. Other calls are a harsh “tchk”, a click, and a guttural “tuc-wuh-h” or “tick cwuhh-h” (Wetmore and Swales 1931, Hardy et al. 1990, Raffaele et al. 1998).

Range and status Hispaniola and offshore islands. Resident. Endemic resident on Hispaniola, including Saona, Gonâve

and Tortue Islands where common (Raffaele et al. 1998).

Habitat and general habits Tropical deciduous forest, tropical lowland evergreen forest, forages in mid-story and canopy; woodland and thicket, bushes on barren mountain slopes, pine forest, shade coffee plantations and gardens; sea level to 1200 m. Inquisitive and fearless (Wetmore and Swales 1931, Faaborg 1985, Stotz et al. 1996,Wunderle and Latta 1996, Raffaele et al. 1998). Occur alone or in mixed-species flocks in pine forests (Latta and Wunderle 1996).They move with long strides along and across branches, creep and crawl through dense branches near the ground, rest quietly for minutes at a time, and fluff feathers in the sun in the early morning (Wetmore and Swales 1931,Wauer 1996, Raffaele et al. 1998).

Food Insects (grasshoppers and mantids, also caterpillars, cockroach and cicadas), small snakes and lizards, including geckos (Wetmore and Swales 1931).

Breeding Season is March to June (Wetmore and Swales 1931, Raffaele et al. 1998). Nest is a flat platform of sticks and leaves, placed on a tree stump. Eggs are dull white or pale blue with a thin white outer layer, 37 ⫻ 25 mm, clutch size 2–3 (Kaempfer 1924, Wetmore and Swales 1931, Stockton de Dod 1981, Raffaele et al. 1998). Incubation and nestling periods are unknown.

364 Thick-billed Cuckoo Pachycoccyx audeberti

Genus Pachycoccyx Cabanis, 1882 Pachycoccyx Cabanis, 1882, Journal für Ornithologie , 30, 230.Type, Cuculus validus Reichenow 1879.The name refers to the broad bill (Gr. pakhus, thick, large; kokkux, cuckoo). Bill broad, culmen strongly curved, nostril slit-like. It is a large cuckoo with no obvious close relationship to other genera. In molecular genetics it is nearly basal to the entire

clade of Old World parasitic cuckoos, or is closer to the koels Eudynamys than to other cuckoos, and the juvenile plumage suggests the pale plumage of juvenile koels in the southern Moluccas and Australia. Only a pelvic skeleton, and no spirit specimens are available in museum collections. One species.

Thick-billed Cuckoo Pachycoccyx audeberti (Schlegel, 1879) Cuculus audeberti (Schlegel, 1879), Notes from the Leyden Museum, 1, 99. (Ambodikilo, Antongil Bay, Madagascar) Polytypic. Three subspecies. Pachycoccyx audeberti audeberti (Schlegel, 1879); Pachycoccyx audeberti validus Reichenow, 1879; Pachycoccyx audeberti brazzae Oustalet, 1886.

Description ADULT: Sexes alike, above uniform slate gray (brown in worn plumage), wing blackish, tail broadly barred brown and black, feathers broad near the tip, T1 black, T2 to T5 with a white tip; underparts unbarred white from chin and throat to belly and under tail coverts, the under tail coverts in some birds with broad black bars; eye-ring yellow, iris brown, bill black above and green below to yellow at base, nostrils slit-like, unlike other parasitic cuckoos except Long-tailed Cuckoo Urodynamis taitensis, feet yellow. JUVENILE: Crown and nape black with white patches and a white collar, back dark brown with broad white spots, wing coverts with large white spots, the feathers of crown, back and wing with white patches at tips are variable among birds, tail feathers narrow, pointed near the tip, under tail coverts unbarred, face mostly white; underparts white; eye-ring yellow, iris dark brown, bill blackish above and grayish green below. NESTLING: Naked at hatching, skin orange; bill yellow, gape orange.The gape darkens with age and is black by fledging, while the palate develops two

large contrasting spots, and the tongue is orange; the bill changes from gray-brown by day eight to black in older nestlings. By 12 days the feathers begin to emerge from quills on wing coverts and flanks, by 14 days all body feathers have begun to emerge and are buffy white, and by 18 days the plumage is buffy white, with a few dark feathers later emerging on the back and with dark brown distal half of the primaries; iris dark brown, eye-ring yellow, bill blackish. SOURCES: AMNH, BMNH, BWYO, CM, FMNH, MCZ, MRAC, RMNH, USNM, ZMB, ZMUC, ZSM (other specimens are listed in Benson and Irwin 1972).

Subspecies Pachycoccyx audeberti audeberti (Schlegel, 1879); larger, lores white; Madagascar; Pachycoccyx audeberti validus (Reichenow, 1879); smaller; Angola, E Zaire to Tanzania, SE Kenya, Malawi, Zambia, N Botswana, Zimbabwe, lores gray; Mozambique, E Transvaal and N Natal; Pachycoccyx audeberti brazzae (Oustalet, 1886); smaller, above black, no white on lores; Sierra Leone to Ghana, Nigeria, Cameroon, Central African Republic, Congo and western Zaire. Birds between northern Lake Tanganyika and the Uele are intermediate between the forms validus and brazzae (Benson and Irwin 1972).

Measurements and weights P. a. validus, Kenya,Tanzania, Zimbabwe, Zambia and Angola; Wing, M (n ⫽ 10) 216–230 (223.3 ⫾ 4.3),

Thick-billed Cuckoo Pachycoccyx audeberti 365 F (n ⫽ 8) 217–228 (220.8 ⫾ 3.3); tail, M 166–198 (183.0 ⫾ 10.9), F 162–193 (181.6 ⫾ 9.9); bill, M 21.9–26.7 (23.4 ⫾ 1.6), F 21.1 ⫾ 23.8 (22.4 ⫾ 1.1); tarsus, M 20.3–25.5 (23.1 ⫾ 1.4), F 22.0–25.0 (22.9 ⫾ 1.0) (AMNH, FMNH, ZMUC, ZSM); P. a. brazzae: Wing, M (n ⫽ 3) 204–232 (218.0), F (n ⫽ 2) 216–233 (224.5), U (n ⫽ 1) 233 (AMNH, Benson and Irwin 1972); P. a. audeberti: Wing, M (n ⫽ 1) 237, F (n ⫽ 1) 238, U (n ⫽ 2) 231, 234 (Benson and Irwin 1972). Weight, M (n ⫽ 1) 92, F (n ⫽ 4) 100–120 (110.5) (Verheyen 1953, Benson and Irwin 1972). Wing formula, P8 ⬎ 7 ⬎ 6 ⬎ 9 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 10 ⬎ 2 ⬎ 1.

Loud whistled “Oui, yes, yes” or “wheee-wheeewheep”, and “weer-wick”, like a goshawk. The whistled call is two or three rising whistles, the first longer and lower than the second, in Ivory Coast, Zaire and Zambia and South Africa (Gillard 1987, Stjernstedt 1993, Chappuis 2000). Birds in Madagascar respond to playback of birds recorded in southern Africa, and the calls in these areas are similar (Goodman and Putnam 1996, S. M. Goodman pers. comm.). Some calls break into an excited shrill babbling (Stevenson and Fanshawe 2002) and another call is a chatter “chee-cher-cher-cher” thought to be of the female (Stjernstedt 1993). Fledglings and courting adults give a two-note squeak (Vernon 1984).

Ghonerezhou Reserve (Irwin 1988, Riddell 1990, 1994). Rare in Madagascar, where unrecorded from 1922 until 1992 when one was seen at Maromizaha forest near Périnet-Analamazaotra, and another was seen in 1993 at Andringitra (Langrand and Sinclair 1994, Goodman and Putnam 1996, Morris and Hawkins 1998). They are non-migratory, or local seasonal migrants. In Liberia, they occur in high forest and logged forest of northern highlands and the southeastern lowlands (Gatter 1997). Resident in Ivory Coast in southern Guinea woodlands in Lamto,Tai and north to Combo (Thiollay 1985). In Ghana and Togo they are rare residents (Grimes 1987, Cheke and Walsh 1996). In Nigeria, near Kagoro they occur in all seasons (Dyer et al. 1986, Elgood et al. 1994). In the Central African Republic they occur in Dzanga-Sangha (AMNH). In southern Sudan, they appear as non-breeding visitors (Nikolaus 1987). In Zambia and Malawi, they appear in nearly all months (Benson and Benson 1977,Aspinwall and Beel 1998), in Zimbabwe they are seen between September and February (Vernon 1967, 1984), and in South Africa most records are in the austral summer (Benson and Irwin 1972, 1973).

Range and status

Habitat and general habits

Africa from Sierra Leone through Zaire and Kenya south to Angola and Transvaal; Madagascar. They are uncommon to rare throughout the species range (Benson and Irwin 1972, 1973), regular in coastal Kenya forest, in Zambia and Zimbabwe in the Zambezi Valley near Lake Kariba, and in southeastern Zimbabwe along the Sabi River and in

Miombo (Brachystegia) woodland and other open woodlands in southern and East Africa, and lowland and riverine forests; in high forest and logged forest in West Africa.The birds are more characteristic of rich miombo (Brachystegia-IsoberliniaJulbernardia woodland than closed-canopy evergreen forest) (Heinrich 1958, Payne 1968, Dyer et al.

Field characters Overall length 36 cm. Large cuckoo with an accipiter hawk-like appearance, unmarked dark above and white below, with a banded tail. The juvenile, white in the nest, is pied with white and black upperparts after it fledges.

Voice

366 Thick-billed Cuckoo Pachycoccyx audeberti 1986, Irwin 1988, Allport and Fanshawe 1994, Gatter 1997). In coastal woodland in the 500⫹ mm rainfall area in East Africa (Lewis and Pomeroy 1989, Zimmerman et al. 1996), in forest in Madagascar (Morris and Hawkins 1998). In the Zambezi Valley and southeastern Zimbabwe they are mainly in the lowlands; in Liberia and in Madagascar they occur to nearly 1000 m.

Food Insects, mainly hairy caterpillars, mopane worms (Imbrasia belina), grasshoppers and mantids (Brooke 1965a, Benson and Irwin 1972, Jones 1996, CM). The cuckoo removes from the nest and eats eggs of its hosts (Vernon 1984).

Displays and breeding behavior Territories are large, several kilometers in length along a river forest (Riddell 1994). In the breeding season cuckoos appear alone, in pairs or small groups. They display in flight with loud calls and slow wingbeats, then one crashes into the top of a tree, perches a few minutes and looks around, then if not joined by another bird it goes on another flight, displaying over several kilometers; the bird sometimes displays in flight along the same route in return. Courtship feeding and copulation occurs when a displaying bird meets another bird. The male may cooperate with the female in parasitizing the host nest (Vernon 1984).

Breeding In Zimbabwe they breed mainly during the late rains but may continue later, September to March, mainly October to November (Vernon 1984), in Zambia they breed in September (Irwin 1988). In Cameroon a female had a developed ovary in March, and a growing juvenile in early July may date from March (Bates 1930, Benson and Irwin 1972). In Liberia they call from April to May and October to November (Gatter 1997).There are few records elsewhere. Brood-parasitic. Hosts, in Africa they parasitize Red-billed Helmet-shrike Prionops (Sigmodus) retzii (Vernon 1984, Irwin 1988), Chestnut-fronted

Helmet-shrike P. scopifrons (Pinto 1959) which is the common “Sigmodus” species in SE Tanzania where these cuckoos occur (ZMUC, ZSM), suspected host Chestnut-bellied Helmet-shrike P. caniceps in Liberia (Gatter 1997) and (once, Zaire) White-crowned Helmet-shrike P. plumata (Louette 1989). In Madagascar, where helmet-shrikes do not occur, the hosts are unknown; Chabert Vanga Leptopterus chabert is suspected (Benson et al. 1976). The cuckoo removes an egg from the nest of its host, sometimes actively knocking the incubating host off the nest; the host chases the cuckoo after the cuckoo takes the egg. The female cuckoo also removes host nestlings from the nest and drops them onto the ground. Eggs are pale blue-green with large brown, gray and lilac blotches mainly in a band around the broad end, 24 ⫻ 17 mm, and closely resemble eggs of the helmet-shrike host species (Vernon 1984, Tarboton 2001). Incubation is c. 13 days (shorter than the host eggs); the egg is often laid a few days after the host has incubated its own clutch. The nestling cuckoo has a broad back with an expanded synsacrum, and it evicts the host eggs and nestlings. It voids foul-smelling feces.The nestling cuckoo differs from the nestling helmetshrikes in the color of the skin and the palate and the cuckoo is not a visual mimic of the host young. Mouth markings of the young cuckoos may have been retained in this basal brood-parasitic cuckoo from the ancestral malkohas.The feathered nestling is white, in contrast to the host nestlings which are gray or brown on the upperparts. Richard Peek and Kit Hustler took the photograph that is reproduced in Color Plate 10. The cuckoo fledges at 84–86 g, nearly twice the mass of an adult helmet-shrike. The nestling period is long, 28–30 days, then after it leaves the nest the young cuckoo is fed up to 50 days. Breeding success, in 29 nests, 46 cuckoo eggs survived to hatch, and 5 of 16 cuckoo chicks survived to fledge. In 29 nests of S. retzii that held a nestling host or cuckoo, 16 had a nestling cuckoo, a 55% rate of parasitism. Levels of parasitism can be high in a local population, with more than half of the nests parasitized over several years and no successful breeding by the host (Vernon 1984).

Dwarf Koel Microdynamis parva 367

Genus Microdynamis Salvadori, 1878 Microdynamis Salvadori, 1878, Annali del Museo Civico di Storia Naturale di Genova 13, p. 461.Type, by subsequent designation, Eudynamis parva Salvadori 1875. The generic name derives from Gr. micros, small; Dynamene, koel, from Gr. δυναµισ, strength, power;

little koel. Nostrils slit-like. Small cuckoo with marked sexual and age dimorphism in plumage, male with face pattern of black cap and black malar streak. No skeletal or alcoholic specimens are available for comparison with other cuckoos. One species.

Dwarf Koel Microdynamis parva (Salvadori, 1875) Eudynamis parva Salvadori, 1875, Annali del Museo Civico di Storia Naturale di Genova 7, p. 986. [?Molucche : ?Tidore (⫽ New Guinea)]. Polytypic. Two subspecies. Microdynamis parva parva (Salvadori, 1875); Microdynamis parva grisescens Mayr and Rand, 1936. Other common names: Black-capped Koel.

Description ADULT: Male, crown and nape glossy black, back brown, wing brown, outer edge of primaries, secondaries and coverts rufous, tail brown, a broad white streak below the eye extends as a buff streak behind the eye, the black behind the eye extends to the shoulder and a glossy black malar streak extends from base of lower bill to side of breast; underparts, throat orange-yellow, breast, belly and under tail coverts unmarked gray or buff, bend of wing unmarked buff to white, inner vane of base of primaries rufous, tip indistinctly darker, no wing bar, under wing coverts light brown; iris red, bill black (under mandible base whitish horn), short and bluntly hooked at the tip, feet blackish gray. Female, has the crown and back uniformly grayish brown, gray streak below eye, throat gray, upper breast gray, breast and flanks gray with faint brown bars; iris hazel to red, bill black, feet blue gray to black. Immature male has the throat streak charcoal gray, the crown brown. JUVENILE: Crown rufous brown, plumage loosewebbed, back brown, rump brown indistinctly barred buff, wing brown, outer vane of primaries with indistinct bars of gray rufous, black, and rufous buff with a black subterminal band and gray-rufous

tip, tail rufous brown; underparts from throat to belly and under tail coverts indistinctly barred light and dark brown, whitish streak below the eye is narrow and indistinct. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, MCZ, MSNG, SMF, UMMZ, ZMB.

Subspecies Microdynamis parva parva (Salvadori, 1875); back rufous brown, underparts buff; lowland forests of southern and eastern New Guinea and highlands of eastern New Guinea, D’Entrecasteaux Islands; Microdynamis parva grisescens Mayr and Rand, 1936; back grayish brown, underparts rufous brown on breast, and the belly and under tail coverts gray; northern New Guinea between Humboldt Bay and Astrolabe Bay. The species was described from a specimen of uncertain locality (“? Molucche : ?Tidore”). Salvadori’s type specimen (MSNG CE-13824) is rufous brown like the birds of southeast New Guinea, and unlike the gray birds of northern New Guinea.The species does not occur in the Moluccas.

Measurements and weights Wing, M (n ⫽ 12) 102–117 (105.4 ⫾ 4.6), F (n ⫽ 8) 95–106 (101.5 ⫾ 4.0); tail, M 88–105 (95.1 ⫾ 6.0), F 81–98 (87.4 ⫾ 5.3); bill, M 16.1–19.8 (18.0 ⫾ 1.0), F 16.7–19.2 (18.0 ⫾ 0.9); tarsus, M 18.6–22.7 (20.5 ⫾ 1.2), F 18.6–20.8 (20.1 ⫾ 0.7) (AMNH, MCZ, SMF, UMMZ, ZMB).

368 Genus Eudynamys Weight, M (n ⫽ 4) 39.3–54 (43.3), F (n ⫽ 2) 49–54 (51.5) (AMNH, Diamond 1972). Wing formula, P8 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⱖ 9 ⬎ 2 ⬎ 1 ⬎ 10. P9 is short in this species.

Field characters Overall length 20 cm. Small brown cuckoo with a short bill and a rounded tail. Male is brown above, has black cap and nape to below the eye, a white line under the eye and a black malar streak; female is grayish brown with a pale gray malar streak and throat. Plumage pattern of the male especially on the face is like that of some female Common Koel Eudynamys scolopacea.

Voice Song is a series of notes of medium to high pitch, series of resonant upslurred whistles, “touei touei touei . . .”, 1/sec, 10 to 30 notes, becoming louder. The whistles sound like a Common Koel but the notes are shorter, faster, not as loud and are higher in pitch. Another call is a rapid, rising, liquid series of notes, the series rising in pitch then leveling at the end, “tewhodohodohodohodohodohodo . . .” (Diamond 1972, Coates 1985, 1991, Beehler et al. 1986, Crouch and Crouch 1993).

Range and status New Guinea and D’Entrecasteaux Islands (Goodenough I, Fergusson I) (Rothschild and Hartert

1907, Ogilvie-Grant 1915, Hartert 1930, Mayr and Rand 1936, Beehler et al. 1986). Resident, widespread, uncommon to fairly common.

Habitat and general habits Forest, forest edge, secondary forest bordering on sago swamp; lowlands to 1450 m, in eastern New Guinea highlands from Mengino, Okasa and Karimui (Diamond 1972, Coates 1985, 1991).

Food Fruit, which is also given in courtship feeding (Rand 1941a, Coates 1985).

Breeding Unknown, presumably brood-parasitic.

Genus Eudynamys Vigors and Horsfield, 1826 Eudynamys Vigors and Horsfield 1826, Transactions of the Linnean Society, London, 15 (pt. 1), p. 303. Cuckoo sexually dimorphic in plumage, males black and females usually streaked brown; juvenile plumage either like the adult or distinct from either sex. Nostrils slit-like. Type, by subsequent designation, Cuculus orientalis Linnaeus (Gray, 1840). Etymology Ευ (“bene”, true), and δυναµις, “L. potentia”, power), referring to the stout bill (“The bill is powerful, the under mandible more particularly . . . ” (Vigors and Horsfield 1826, pp. 303–4). R. P. Lesson, 1828, Manuel d’Ornithologie, vol. 2, p. 123, emended the spelling to Eudynamis Vigors and Horsfield. Greek

δυναµις is feminine in gender, a regular third declension noun. Although Greek “ι” (iota) in current convention is transliterated as “i” not “y”, ICZN (1999) article 32.5.1.1 states that an incorrect transliteration is not an incorrect original spelling: for this reason the original spelling Eudynamys by Vigors and Horsfield is correct and the gender is feminine. Vigors and Horsfield (1826) included two species, E. orientalis (Cuculus orientalis of Linnaeus 1766) and E. Flindersii (a manuscript name of Latham). One species (more proposed by some authors), widespread from India and eastern Asia through Indonesia, the Bismarck Archipelago, Solomon Islands and Australia.

Common Koel Eudynamys scolopacea 369

Common Koel Eudynamys scolopacea (Linnaeus, 1758) Cuculus scolopaceus Linnaeus, 1758, Systema Naturae, ed. 10, 1, p. 168. (Malabar) Polytypic. Fifteen subspecies. Eudynamys scolopacea scolopacea (Linnaeus, 1758); Eudynamys scolopacea orientalis (Linnaeus, 1766); Eudynamys scolopacea mindanensis (Linnaeus, 1766); Eudynamys scolopacea cyanocephala (Latham 1801); Eudynamys scolopacea rufiventer (Lesson, 1830); Eudynamis melanorhyncha S. Müller, 1843; Eudynamys scolopacea chinensis Cabanis and Heine, 1863; Eudynamys scolopacea malayana Cabanis and Heine, 1863; Eudynamys scolopacea salvadorii Hartert, 1900; Eudynamys scolopacea everetti Hartert 1900; Eudynamys scolopacea alberti Rothschild and Hartert, 1907; Eudynamys scolopacea minima van Oordt, 1911; Eudynamys scolopacea subcyanocephala Mathews, 1912; Eudynamys scolopacea harterti Ingram, 1912; Eudynamys scolopacea hybrida Diamond, 2002. Other names: Asian Koel (Asian taxa); Celebes Koel, Sulawesi Koel, Black-billed Koel (Sulawesi form); Australian Koel, Blue-headed Koel, Cooee, Rain Bird (Australian forms).

Description ADULT: (E. s. scolopacea) Male, glossy black; iris red, bill light green, feet gray. Female, crown brown with rufous streaks, back, rump and wing coverts dark brown with white and buff spots, wing dark brown with white and buff bars, tail dark brown crossed with thin white to rufous bars; underparts whitish, throat and breast streaked blackish brown, breast spotted blackish brown, and belly and under tail coverts barred blackish brown, under wing coverts whitish with thin blackish bars; iris light brown to orange red. JUVENILE: (E. s. scolopacea) Male dull slate with buff tips on feathers of breast, belly and wing (some feathers also barred buff), female darker above than adult female with blackish head and breast; iris brown, bill black. The rectrices are narrower and more pointed than in the adult. Bird molts directly from buff-tipped slate juvenile plumage into adult, the mouth lining unmarked, bright orangish pink, edge of bill and gape yellow, turning black soon after fledging.

NESTLING: In India, naked, skin brownish (vs. pinkish in crows) changing in a few days to black with a gray tinge. SOURCES: AMNH, BMNH, FMNH, MCZ, MSNG, MVZ, RMNH, ROM, SMF, SMTD, UMMZ, USNM, UWBM,YPM, ZMB.

Subspecies Eudynamys scolopacea scolopacea (Linnaeus, 1758); as above, female head and back dark brown with whitish streaks and spots; Pakistan, India, Nepal, Sri Lanka, Laccadives, Maldives; Eudynamys scolopacea chinensis (Cabanis and Heine, 1863); female with no rufous on crown, larger; southern China, continental Indochina; Eudynamys scolopacea harterti (Ingram, 1912); smaller, bill longer than chinensis; Hainan; Eudynamys scolopacea malayana (Cabanis and Heine, 1863); female crown streaks, dorsal spots and underparts washed rufous; S Burma,Thailand, Malay Peninsula, Sumatra, Bangka, Lesser Sundas (Lombok, Sumbawa, Satonda, ?Komodo, Flores, Besar, Paloe), Borneo; Eudynamys scolopacea mindanensis (Linnaeus, 1766) (includes E. s. paraguena (Hachisuka, 1934), from Palawan, and E. s. corvina (Stresemann, 1931), from Halmahera); light areas rufous and the dark ventral bars narrower than malayana; the Philippines (including Palawan and Babuyanes Islands), islands NE of Sulawesi (Talaud Islands (Karakelong, Lirung), Sangihe, Siau, Ruang, Manterawu); northern Moluccas (Morotai, Halmahera, Ternate, Tidore, Moti, Bacan); Eudynamys scolopacea rufiventer (Lesson, 1830); bill yellowish horn, female heavily spotted and barred rufous and black above, head and throat mainly rufous, pale streak below the eye, underparts buff with fine dark barring; New Guinea (except southern Irian Jaya); Eudynamys scolopacea minima van Oordt 1911; small, female crown black, back with small rufous and white spots, tail rufous and dark brown bars, face black with pale rufous streak below the eye, black sides of throat and darker rufous sides of

370 Common Koel Eudynamys scolopacea upper breast, underparts unbarred light rufous; southwestern New Guinea; Eudynamys scolopacea salvadorii Hartert, 1900; female similar to rufiventer, whitish spots on wing coverts, breast and belly buffy white with narrow black bars, under wing coverts dark buff with fine black bars, larger than rufiventer; Bismarck Archipelago; Eudynamys scolopacea hybrida Diamond, 2000; female belly intermediate in color between E. s. rufiventer and E. s. salvadorii, female tail pattern like E. s. salvadorii, both sexes like E. s. rufiventer in size; Long Island, between New Guinea and New Britain; Eudynamys scolopacea alberti Rothschild and Hartert, 1907; female barred rufous and dark brown above, wing coverts with no white spots, face dark with thin rufous streak below eye, throat rufous with black streaks, breast and belly rufous (not whitish or buff) and finely barred (rufous bars 5–6 mm, black bars 0.5 m), under wing coverts rufous with fine black bars, smaller than salvadorii; juvenile, above barred rufous and dark brown from head to tail (rufous bars 4–5 mm, dark bars 2 mm), below rufous with dark brown bars (rufous bars 4–5 mm, dark bars 1(2) mm), tail rufous and dark brown (rufous bars 6–7 mm, dark bars 3–7 mm); Solomon Islands; Eudynamys scolopacrea melanorhyncha S. Müller, 1843; male black, glossed blue-violet, with a black bill; female bill black to horn-color, plumage polymorphic: (1) a rare morph black like male but the gloss blue, (2) crown and face black with a bluegreen gloss, upperparts, wing and tail blackish with a bronze gloss, malar streak white malar, underparts with the throat and breast blackish or gray, breast and belly buff to rusty fawn with indistinct black bars, and (3) crown and face black (crown in some with rusty streaks), back, wings, rump and tail barred rufous and black, malar streak white, throat rusty with black streaks, breast to belly and under tail coverts buff with narrow blackish bars (all three forms taken in laying condition, Stresemann 1940); bill black to horn color; the juveniles are sooty black; Sulawesi, Banggai, Muna, Togian Islands, Peleng and Sula Islands (Taliabu, Seho); Eudynamys scolopacea orientalis (Linnaeus, 1766) (includes E. s. picata S. Müller, 1843); female crown

black to rufous with black streaks, back with fewer and larger rufous spots and bars than in northern birds, throat black or with black patches, breast and belly rich buffy rufous with little or no barring, bill brown or greenish white; juvenile rufous buff with few bars or spots; C and S Moluccas (Buru, Manipa, Kelang, Seram, Ambon, Tujuh, Watubela Islands); Eudynamys scolopacea everetti Hartert 1900; female variable (crown and nape black, lower back with spots and bars, wing brownish, tail rufous with dark brown bars, buff streak below eye, throat black, or variegated black and ochre, or buff with broad black malar streak, underparts dark buff with irregular thick black bars (not white as in cyanocephala and subcyanocephala, which have narrow and regular bars), juvenile buffy and rusty above, black crown patch and black ear patch, underparts light buffy with thin brown bars, paler than adult female; Sumba to Timor and Roma, Kai Islands; Eudynamys scolopacea cyanocephala (Latham 1801); male, glossy black; iris red, bill blue-gray to horn; female, crown black with fine white spots and streaks, back and wing dark brown with fine white spots and bars, tail dark brown with narrow white bars, face black, white streak below eye extends to breast, broad black malar streak, throat buff, breast and belly whitish with fine blackish bars; iris red, bill light green; juvenile, crown rufous buff, above buff barred dark brown, face buff with black ear streak and malar streak, throat buff, breast and belly white to buff with fine dark bars; iris brown, bill whitish to light greenish to black; Torres Strait islands north to Boigu and Darnley, N and E Queensland, west to the lower Norman River and north to Coen and Cape York and islands off the east coast as far as the Capricorn group, and in New South Wales; Eudynamys scolopacea subcyanocephala Mathews, 1912; female crown black; northern Australia (Western Australia, Northern Territories, western Queensland south to Mt Isa and Dolomote and east to the Cloncurry); migrant to New Guinea.

Systematic notes Juveniles from Asia south through the northern Moluccas have dark plumage, juveniles from New

Common Koel Eudynamys scolopacea 371 Guinea (rufiventer) to the Solomon Islands (alberti) are dark, and southern juveniles from the central Moluccas to Australia are pale and buffy and these southern birds are sometimes considered a separate species E. orientalis. Koels of Sulawesi are sometimes considered a distinct species, E. melanorhyncha. Wing shape and female plumage vary among geographic regions, but not in a concordant pattern. The mtDNA sequences indicate that melanorhyncha is closely related to alberti and cyanocephala of the Solomon Islands and Australia and to E. s. scolopacea of the Asian continent. Birds on the Bay of Bengal islands (Andamans and Nicobars) (Ripley and Beehler 1989) were described as a subspecies E. s. dolosa Ripley, 1946, with small size and large rufous spots in female plumage. On these islands, females vary from rufous to no rufous on head, perhaps from repeated immigration of E. s. scolopacea from India (where they have little rufous) and E. s. malayana from the Malay Peninsula (where they are quite rufous), or nonbreeding migrants from the mainland; specimen records are from winter. On Simeulue I west of northern Sumatra, E. s. simalurensis Junge 1936 was described as smaller than malayana and the females with larger rufous spots above. Females in the type series vary in spots and color, and the size of males and females falls within the range of E. s. malayana (type series of E. s. simalurensis, wing, M (n ⫽ 11) 186–197 (194.1 ⫾ 3.6), F (n ⫽ 6) 189–200 (194.8 ⫾ 4.2) (Junge 1936, RMNH). In the Philippines, McGregor (1904, 1909) found Koels of variable size on Calayan and Fuga in the Babuyanes Islands north of Luzon, and recognized two kinds, small birds called E. mindanensis and large birds that he described as E. frater.Another male taken in 1895 in Fuga by Whitehead (AMNH 627271, wing 217) was near the mean size of males on Luzon.The type series of frater (wing, M (n ⫽ 8) 231–260 (241.0 ⫾ 8.8), F (n ⫽ 1) 240 (McGregor 1904)) and the other Calayan Koels were collected in 1903. McGregor’s other birds on Calayan, which he referred to mindanensis, were smaller (wing, M (n ⫽ 5) 184–201 (192.2 ⫾ 7.8). Four birds identified as frater and taken by McGregor in 1903 are in FMNH and SMTD (others were destroyed in Manila during the war, and the two type specimens

designated by McGregor have not been seen). In others in the type series, male SMTD C-20589 was taken on 15 Sept and has a few white feathers and a wing of 273, matching the largest bird in McGregor’s list at 10.25 “. Male FMNH 20215 was taken on 12 Oct, the wing 243 matches the 9.2” reported for male no. 3612 collected on that date and listed by McGregor (1904). The others are females (FMNH 19976, 11 Nov, wing 215; FMNH 19978, 22 Dec, wing 226), neither listed by McGregor, who mentioned only one female (no. 3201), taken 18 Sept, dark with feather shafts and spots of rufous above and the underparts buff barred with black. The first is similar to McGregor’s description and looks like females from Bohol, and the second is more blackish; while in size the first is intermediate between the small and large birds in McGregor’s list, and the second is like the smallest bird in the type series of frater. The largest bird in McGregor’s list is larger than all other Koels, and the range of size is extreme. Birds taken on Fuga in March,April and August in 1979 and 1980 on later DMNH duPont expeditions had male wings 185, 204, 218 and 220 and female wings 204, 206, 217, the last one larger than most birds in the Philippines. There is no evidence that two populations breed on Fuga (both small and large birds were in molt), and specimens from Fuga other than those listed by McGregor are intermediate in size between the classes reported by McGregor. E. s. corvina Stresemann 1931 from the northern Moluccas was described as like E. s. mindanensis in female plumage but larger. My measurements of birds from Halmahera and nearby islands (Ternate, Obi, Siku) overlap those of E. s. mindanensis from Mindanao, and one bird is smaller (MSNG 13786, M, wing 197) than the mean of males from Mindanao. In measurements corvina is not consistently distinct from mindanensis. E. s. melanorhyncha, sometimes considered a separate species, differs in bill color. The use of noncrow hosts suggested to White and Bruce (1986) it was a separate species; nevertheless, some Asian populations of Common Koel parasitize other hosts, as the mynahs in Thailand, the Malay Peninsula and Palawan.The wing is rounded as in some other tropical populations of Common Koel.The nestlings are

372 Common Koel Eudynamys scolopacea undescribed. No geographic overlap is known between E. s. melanorhyncha and other koels. East of Sulawesi, in the Sula Islands (Taliabu) where koels have black bills and a few males have white on the head or even on the body, the population is sometimes recognized as E. s. facialis Wallace 1862 (Hartert 1900, Eck 1976,White and Bruce 1986). Not all males in the population have white on the head (Wallace 1862, Eck 1977). Although birds in the Sula Islands average smaller than birds on Sulawesi, they overlap in size (BMNH, SMTD). In New Guinea the females vary in the extent of rufous and black on the head, as noted by Hartert et al. (1936) and this variation led to the idea that both rufiventer and subcyanocephala occur in the same region (Rand 1941a); however, subcyanocephala may only winter there (Mayr 1937). Similarly both orientalis and subcyanocephala occur on Buru and Seram where subcyanocephala may be nonbreeding visitors (Bowler and Taylor 1989; MSNG, SMTD). Koels on Long, Tolokiwa and Crown Islands northeast of New Guinea and west of the Bismarck Archipelago are distinct and have been described as a subspecies, E. s. hybrida. Birds vary little from each other in appearance, small like New Guinea birds E. s. rufiventer, and female plumage has rufous underparts nearly as dark as E. s. salvadorii and white wing spots like E. s. salvadorii of New Britain (AMNH). The birds may be a hybrid population that formed during the past three centuries, after a volcano defaunated the islands, and birds from New Guinea and New Britain recolonized the islands (Mayr and Diamond 2001, Diamond 2002). A female in similar plumage but larger was taken on another volcanic island, Umboi (⫽ Rook I, Rothschild and Hartert 1914b,AMNH 627379); Umboi was not involved in the same volcanic episode (Mayr and Diamond 2001). In southwestern New Guinea, small koels E. s. minima occur in the lowland rivers region (Mimika, Setakwa (Otakwa, Utakwa), Lorentz (Noord) and Digul (Digoel) Rivers). No koel specimens seem to be known further east in the lowlands near Merauke and Koembe (Mees 1982).The female (MZB 17852, taken 20 April 1913 on the Digul River at 140 m) is rufous (but not as dark as the northern New Guinea females) and has a

blackish crown. In plumage pattern and color the females are intermediate between New Guinea E. s. rufiventer and the Australian koels.

Measurements and weights Eudynamys scolopacea scolopacea, India: Wing, M (n ⫽ 10) 187–209 (194.8 ⫾ 7.0), F (n ⫽ 17) 172–195 (188.6 ⫾ 4.4); tail, M 170–205 (188.6 ⫾ 12.7), F 175–200 (185.2 ⫾ 7.5); bill, M 26–32 (28.3 ⫾ 2.1), F 26–30 (27.6 ⫾ 1.2); tarsus, M 31–36 (33.4 ⫾ 1.8), F 30–35 (32.8 ⫾ 1.6) (UMMZ); E. s. chinensis, China,Vietnam and Laos: Wing, M (n ⫽ 14) 192–213 (203.0 ⫾ 7.4), F (n ⫽ 14) 187–227 (201.9 ⫾ 12.4); bill, M 25–30 (26.8 ⫾ 1.4), F 23–28 (26.0 ⫾ 1.5) (AMNH, FMNH, UMMZ, USNM); E. s. harterti, Hainan: Wing, M (n ⫽ 13) 199–220 (206.9 ⫾ 5.5), F (n ⫽ 12) 196–217 (204.4 ⫾ 8.1); bill, M 25–32 (29.0 ⫾ 1.6), F 25–31 (28.1 ⫾ 2.0) (AMNH, SMF); E. s. malayana,Thailand: Wing, M (n ⫽ 12) 199– 238 (213.0 ⫾ 11.7), F (n ⫽ 9) 195–213 (205.0 ⫾ 5.8) (USNM); E. s. mindanensis, Palawan: Wing, M (n ⫽ 16) 177–217 (198.4 ⫾ 6.8), F (n ⫽ 3) 193–201 (199.3) (AMNH, SMF, USNM, YPM); Luzon: Wing, M (n ⫽ 23) 191–215 (199.8 ⫾ 6.2), F (n ⫽ 6) 185–203 (195.7 ⫾ 8.4); tarsus, M 30.4–37.0 (34.1 ⫾ 1.9), F 30.2–35.6 (32.5 ⫾ 1.6) (AMNH, FMNH, SMF, USNM); Calayan and Fuga: Wing, M (n ⫽ 10) 196–273 (234.3 ⫾ 21.5), F (n ⫽ 4) 204–226 (215.5 ⫾ 9.0) (AMNH, DMNH, FMNH, SMTD, USNM, YPM); Mindanao: Wing, M (n ⫽ 9) 192–204 (197.6 ⫾ 5.8), F (n ⫽ 6) 187–201 (194.5 ⫾ 6.0) (FMNH, SMF, USNM); Halmahera, Obi, Siku and Ternate: Wing, M (n ⫽ 7) 197–215 (208.5 ⫾ 5.4), F (n ⫽ 2) 189–214 (201.5) (AMNH, MSNG, MZB, SMF, USNM, ZMB); E. s. rufiventer: Wing, M (n ⫽ 10) 184–200 (191.8 ⫾ 4.7), F (n ⫽ 12) 181–198 (188.5 ⫾ 5.0) (AMNH, SMF); E. s. minima: Wing, M (n ⫽ 2) 168–178 (173), F (n ⫽ 1) 170 (MZB, RMNH); E. s. salvadorii: Wing, M (n ⫽ 14) 203–217 (199.2 ⫾ 4.4), F (n ⫽ 9) 193–210 (202.4 ⫾ 7.8) (AMNH, MCZ); E. s. hybrida: Wing, M (n ⫽ 14) 184–201 (189.5 ⫾ 4.4), F (n ⫽ 5) 180–196 (187.2 ⫾ 6.7) (AMNH);

Common Koel Eudynamys scolopacea 373 E. s. alberti: Wing, M (n ⫽ 12) 167–201 (183.3 ⫾ 8.5), F (n ⫽ 11) 165–184 (175.2 ⫾ 5.9) (AMNH, BMNH, MCZ, USNM); E. s. melanorhyncha: Wing, M (n ⫽ 19) 187–211 (203.9 ⫾ 4.9), F (n ⫽ 11) 180–213 (190.5 ⫾ 6.6); tail, M 173–217 (196.2 ⫾ 12.2), F 188–209 (194.5 ⫾ 6.7); bill, M 25–35 (28.9 ⫾ 3.1), F 25–32 (18.2 ⫾ 2.5); tarsus, M 28–36 (31.8 ⫾ 2.3), F 30–34 (31.3 ⫾ 1.4) (MCZ, MSNG, UMMZ, USNM, ZMS); E. s. orientalis: Wing, M (n ⫽ 17) 201–232 (220.8 ⫾ 7.7), F (n ⫽ 13) 209–229 (217.5 ⫾ 6.7) (AMNH, RMNH); E. s. everetti: Wing, M (September–December) (n ⫽ 6) 190–217 (204.7 ⫾ 5.5), F (n ⫽ 6) 193–205 (197.0 ⫾ 6.6) (AMNH); E. s. cyanocephala: Wing, M (n ⫽ 23) 204–230 (217.2 ⫾ 6.9), F (n ⫽ 32) 200–225 (214.6 ⫾ 6.7); tarsus, M 31.5–38.0 (35.5 ⫾ 1.8), F 33.3–38.5 (35.9 ⫾ 1.3) (E Australia south of 22°S, Higgins 1999); E. s. subcyanocephala: Wing, M (n ⫽ 22) 193–221 (212.3 ⫾ 6.6), F (n ⫽ 9) 193–214 (204.9 ⫾ 6.2) (NW Queensland, Northern Territories and Western Australia, Higgins 1999). Weight, E. s. scolopacea: F (n ⫽ 1) 229 (Ali and Ripley 1969); E. s. malayana: M (n ⫽ 2) 202–242 (222) (Wells 1999); E. s. melanorhyncha: M (n ⫽ 9) 133–231.1 (202.1), F (n ⫽ 4) 191.5–244 (217.1) (FMNH); E. s. cyanocephala: M (n ⫽ 16) 175–340 (261.9), F (n ⫽ 23) 167.3–330 (257.0); E. s. subcyanocephala: M (n ⫽ 8) 135–295 (199.8), F (n ⫽ 3) 169–290 (216.0) (Higgins 1999); E. s. rufiventer: M (n ⫽ 13) 188–254 (219), F (n ⫽ 5) 202–234 (219), E. s. salvadorii: M (n ⫽ 5) 205–330 (275), E. s. hybrida: M (n ⫽ 8) 152–198 (173) (Diamond 2002); E. s. alberti: M (n ⫽ 2) 143–190 (167), F (n ⫽ 4) 148–184 (165) (AMNH, BMNH). Wing formula, E. s. scolopacea: P8 ⬎ 7 ⬎ 6 ⬎ 9 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10. The wing shape varies regionally—northern and southern birds have pointed wings and tropical birds have rounded wings. Migrant subspecies in the more temperate north and south and resident New Guinea birds have wing P8 ⬎ 7 ⬎ 6 ⬎ 9 ⱖ 5 (India scolopacea, China chinensis, Hainan harterti,Thailand and Malay Peninsula malayana, Australia cyanocephala and sub-

cyanocephala, and New Guinea rufiventer), whereas other resident tropical forms have wing P7 ⱖ 6 ⬎ 8 ⬎ 5 ⬎ 9 (Sulawesi melanorhyncha, Halmahera and the Philippines mindanensis); or P7 ⬎ 8 ⬎ 6 ⬎ 9 ⱖ 5 (Bismarck Archipelago salvadorii, Solomon Islands alberti).

Field characters Overall length 42–44 cm. Large arboreal cuckoo with a long rounded tail and a pale green bill (black in Sulawesi), the male all black, and the female brown and buffy with spots and streaks of buff or white and varying in appearance from region to region. The flight is fast, strong and direct.

Voice Call loudly, often and early, well before dawn. In Asia, territorial call of male is a loud whistled shriek,“ko-el”,“ke-woo”,“who are you” or “you’re ILL”, with stress on second note (Hume and Davison 1878, Smythies 1940, Ali and Ripley 1969, King and Dickinson 1975, Subrahmanyam and Krishnamoorthy 1981, Lekagul and Round 1991, Scharringa 1999, Kennedy et al. 2000, Wells 1999, Robson 2000a, Sheldon et al. 2001, Supari 2003). The loud mellow “ko-el” notes, each note 0.5 sec, are repeated once in 2 sec, and a shorter note is given with accelerating soft “vyuk vyuk” phrases (Martens and Eck 1995). Calls a series of 5–10 phrases rising in tempo, pitch and loudness, day and night. Other calls are a “kwow kwow-kwowkwow” excited bubbling. Female gives a shrill “kikkik-kik-kik”. In the New Guinea region on Karkar and Bagabag Islands the calls are (1) series of slurred notes rising in pitch “coooei” or “coooel” and (2) rapid series of notes given in pairs, rising then level in pitch, “duli-duli . . .”, the alarm is a staccato “week week week WEEK . . .”, females give a “piip piip piip piip” (Diamond 1972, Diamond and LeCroy 1979, Coates 1985, 2001, Beehler et al. 1986). In Bougainville and the Solomon Islands, calls include two loud whistles, the second one higher, ‘coo-ee”, the phrases becoming louder

374 Common Koel Eudynamys scolopacea (Cain and Galbraith 1954, Hadden 1981, Doughty et al. 1999). In Sulawesi, a mournful modulated whistle, fast, rising and falling, “wrr-wrr-wrr-wrr-wrr-wrr”, given by the male in the forest canopy and often at night and before dawn (Meyer 1879, Holmes and Phillipps 1996, NSA), is perhaps the call described as “oeoe-oeoeoeoe”, the “oe” like a flute (Coomans de Ruiter and Maurenbrecher 1948) and it is like the “wurroo” of other populations as in Australia. Other calls resemble those of koels elsewhere, including (1) a melancholy whistle “kuOw” or “whooo” in a series of 2–5 similar notes, each series beginning low and becoming increasingly louder, higher in pitch and more rapid towards the end (5 notes in 10 sec), and (2) a crescendo of excited bubbling calls from both sexes (Stresemann 1940, Holmes and Phillipps 1996, Coates and Bishop 1997, Riley 1997). In Australia, calls vary with behavior context and the season. In southern Queensland, males give (1) a loud “cooee” (or “kooeel”), (2) “wurroo” in ascending series, (3) single high-pitched “whik” at the beginning or end of a bout of “wurroo” calls, and (4) a soft clucking “wuk-wuk-wuk” when courting a female; while females give (5) a high piping screech “keek” in rapid bouts,“keek-keek-keek-keek”. Calls apparently “wurroo” have also been described as rising “quoy-quoy-quoy-quoy” and a falsetto “quodelquodel-quodel”. Early in the season most calls are “cooee”, then from December to February when the hosts are nesting most calls are “wurroo”. Calling rates are highest at dawn and dusk with c. 150 calls an hour, more than twice the rate heard in daytime, and birds also call at night. Males call from certain trees, often fig trees with dense foliage, a wide canopy, and taller than trees where they do not call. Male and female often call in a regular pattern, “keek” by the female followed by “wurroo” by the male, with the duet usually begun by the female (Maller and Jones 2001). These different calls have also been recorded elsewhere in New South Wales (Buckingham and Jackson 1990). In Western Australia, Koel calls have a different pattern (Hutchinson 1983). Calls are similar across the range of Common Koel, with the “ko-el” and “vyuk vyuk” of India and the “cooee” and “wurroo” calls of Queensland

nearly the same in audiospectrograms. Reports of regional differences may confuse the variety of calls given in different contexts and seasons. Field recordings are needed in other regions to compare with the calls in Queensland, where they have been described in most detail. Begging calls of young after fledging are loud trills and squeaks, “wheeet-oop-wheeet-wheeetwheet-op”, interspersed with high-pitched screeches, and no apparent mimicry of begging calls of the hosts’ young; the young koels are louder than young host Figbirds Sphecotheres viridis (Gosper 1997, Maller and Jones 2001).

Range and status Asia from Pakistan (mainly Punjab and lower Indus regions to Sind), India and southern China through southeast Asia, the Malay Peninsula, the Philippines, Indonesia, New Guinea and nearshore islands, and the Bismarcks (Crown I to New Ireland) to the Solomon Islands and Australia. Resident or irregular movements. Local migrant and nomadic in northern and southern temperate regions, resident in the tropics (Stevens 1915, Phillips 1948, Traylor 1967, Zacharias and Gaston 1983, Guangdong et al. 1983, Inskipp and Inskipp 1985, Cheng 1991, Roberts 1991, Kotagama and Fernando 1994, McClure 1998, Grimmett et al. 1999,Thomas and Poole 2003). In India, this is the most common cuckoo around Poona, Maharashtra (Mahabal and Lamba 1987).The

Common Koel Eudynamys scolopacea 375 species range expanded northward in Sind when open irrigation canals were built in the early 20th century (Eates 1937, Roberts 1991). In Burma, in the Shan States E. s. chinensis occur from April to June and disappear by July (Osmaston 1916), in the Irrawaddy Plains they occur both in cold weather and in the breeding season (Macdonald 1906), in southern Burma they occur from March to July and are absent from August to November (Smythies 1940, 1986). Koels have appeared in Abu Dhabi and Dubai in February and March (Richardson 1990) and in Oman they have appeared in March and April and in October and November (Gallagher and Woodcock 1980, Eriksen 1991). An October specimen from NE Afghanistan (Bagrami, AMNH 465324) may also be a migrant. Migrants appear in passage in Tonkin (Robson 2000a), and birds winter on coastal islands off the Malay Peninsula (Wells 1999). Records on the Maldives and Lakshadweep are of unknown seasonality. In the Malay Peninsula where they were coastal and occurred mainly as migrants, koels bred only along the northern east coastal plains, halfway down the peninsula, parasitizing Jungle Crows Corvus megarhynchos (Chasen 1939, Medway and Wells 1976). Later they appeared in the Penang area where they used House Crows Corvus splendens, which had spread there from where introduced at Klang in 1903; koels parasitized House Crows at Penang by the early 1980s.The koels’ advance also coincided with the spread of Common Mynahs Acridotheres tristis into cleared habitats. In the east, Koels still use Jungle Crow, but in the west and central parts of the peninsula they use mainly the House Crow. Koels first appeared as residents in Singapore in the 1980s; they were listed as uncommon winter visitors in 1987, but in the first half of the decade beginning in 2000 they have become one of the five most conspicuous birds. Koels completed “a remarkable case of rapid spread . . . in ten years” when their range expanded inland as House Crow expanded their own breeding distribution (Medway and Wells 1976, Hails 1987, Davison and Fook 1995, Wells 1999, Lim Kim Seng 2003). In Borneo, Koels winter from September to April; some may be resident and breed (Smythies 1999, Sheldon et al. 2001).

Australian Koels E. s. cubcyanocephala and cyanocephala arrive in northern Australia in August to October. In the east they occur southward to about 36° S from September to November where they breed in summer and depart in January to April; they have occurred in South Australia (Allen 2003). A few may stay the winter in Queensland (Higgins 1999). E. s. subcyanocephala winter in the Moluccas north to Buru (AMNH 627333), Lesser Sundas, Timor and Tanimbar, and both Australian forms winter in New Guinea and the Aru Islands (Coates 1991). A Koel banded in New South Wales was recovered in New Guinea, a distance of 2950 km from its banding site (Higgins 1999) and gives direct evidence of movement between Australia and New Guinea. Common in much of their range, koels are conspicuous when calling, otherwise they are shy and easily overlooked.Widespread in the austral winter and occur on remote oceanic islands (Salvadori 1880).

Habitat and general habits Open lowland forest, forest canopy, lightly wooded areas mainly secondary forest and forest edge, remnant woodlands where large trees remain, monsoon forest, edge and scrub, riverine scrub, oil palm, cocoa, heath forest, orchards, gardens, in towns near human habitation, around Ficus figs and other fruiting trees, and mangroves, in tree crop plantations in coastal lowlands.They occur mainly in the lowlands. In dry areas they occur in areas of irrigation with orchard and tree plantations. In India they live from the coastal plains to 1000 m in the peninsular hills and to 1800 m in the Himalayas and Peninsular ranges (Grubh and Ali 1974, Price 1979, Roberts 1991, Martens and Eck 1995, Grimmett et al. 1999).They occur from sea level to 1200 m in southeast Asia (King and Dickinson 1975, Wells 1999), they are mainly coastal in the Philippines (Rabor 1977) and Kalimantan (Smythies 1981), they are mainly coastal and rarely occur to 1200 m in Sabah (Sheldon et al. 2001); in Sulawesi from the coast to 1500 m (Coates and Bishop 1997), in Taliabu, Sula Islands, to 500 m (Stone et al. 1997), in the Sangihe lowlands to 800 m and in Flores to 1200 m, rare above 350 m

376 Common Koel Eudynamys scolopacea (Stresemann 1940, Rozendaal and Dekker 1989, Coates and Bishop 1997, Riley 1997). In New Guinea they occur in lowlands and hills locally to 1500 m (Coates 2001), while in the Solomon Islands they are in primary lowland forest (Blaber 1990, Kratter et al. 2001). In Australia they live from sea level to 1000 m, in wooded habitats of monsoon forest, vine thickets, wet broad-leafed sclerophyll forest, and open forest and woodland, in high trees and around fruiting trees, in towns and suburban gardens and parks with figs, mango and mistletoe (Higgins 1999). Arboreal, they feed in tree canopy on fruiting trees. In Sri Lanka,“This is, without exception, the most noisy bird in the country” (Lewis 1898).

Food Fruit, especially figs Ficus, mulberries Morus, buffalo thorn Zizyphus, papaya Carica papaya, guavas, berries, Brazilian cherries Eugenia uniflora, wild caper Capparis sepiaria and tamarinds Tamarindus indica, star-chestnut Sterculia foetida, and fruit of other trees and shrubs including Zizyphus oenoplia, rose-apple Memecylon umbellatum, oleander Thevetia peruviana, macaranga Macaranga peltata, capers Piper nigrum, and Santalum album, and nuts of oil palm Elaeis guineensis and other palms Archontophoenix alexandrae. They take fruits up to 41 mm in size and spit out the large seeds ( Jerdon 1862). Fruit, particularly nutmegs which it swallows whole, also other fruits such as banyan figs (Meyer 1879, Watling 1983); the koel is a dispersal agent of this fruit. Koels take a few insects (grasshoppers, mantids, stick insects, caterpillars), snails and flower nectar (Smythies 1940, Krishnan 1952, Vernon 1968, Hoogerwerf 1969, Rajasingh and Rajasingh 1970, Gaston 1981, Gonzales 1983, Zacharias and Gaston 1983, Crouther 1985, Sody 1989, Midya and Brahmachary 1991, Goodman et al. 1995, Higgins 1999, Sheldon et al. 2001). Females eat eggs— one took four from a bulbul nest in about 4 sec, then flew, and had bits of 3 eggs and an intact egg in her crop (Ali 1931). Koels also take small birds (Higgins 1999). Koels also take nectar of flowering trees Erythrina indica (Saha and Dasgupta 1992). Nestling Koels are reared not only by insect-eating hosts but also by Figbirds Sphecotheres viridis, which

feed both regurgitated fruit and occasional insects to the young.

Displays and breeding behavior In Asia, Koels occur in pairs, and both sexes also mate with other individuals (Dharmakumarsinhji 1954, Lamba 1969). In some areas the males hold territories as close as 200 m to each other and they sing in habitat where no hosts occur, as on some islands, suggesting dispersed arenas of males which females visit to mate, then lay elsewhere (Wells 1999), or aggregations of social fruit-eating birds as in parrots and doves.The male chases the female in flight, both sexes call, and the female perches in a tree. Birds occur in pairs in Australia. In courtship, the male flies to the female, hops after her, droops the wings, spreads and lowers the tail and bows low, repeating this when the female stands still; the male then calls loudly. In copulation, the female flies to the calling male, utters a short ‘wuk’ call, the male approaches her and mounts (Crowther 1985, Roberts 1991, Higgins 1999). In courtship feeding, the male feeds fruit to the female (Lamba 1969, Higgins 1999, Wells 1999). The male may distract the crow from its nest and allow the female to lay in the nest, although she usually visits the nest alone (Lamba 1969, 1976).

Breeding and life cycle In Pakistan they lay from May to August (Roberts 1991), in northern India from March to August (mainly May–June) (Inglis 1903, Ali and Ripley 1969, Lamba 1976), in Bengal in June and July, in Dacca in December and January and again from May to July, in the eastern Ghats they call in March and April, in southern India they breed in December and January in the cool season, although crows nest from September to May and Koels are present all year (Baker 1934, Ali 1953, Price 1979, Zacharias and Gaston 1983); in Sri Lanka they breed from April to August (Ali and Ripley 1969). In Burma on the Irrawaddy Plains they breed in May and June (Macdonald 1906); in the Malay Peninsula they lay from February to April and they sing from December to July and sometimes as late as September (Wells 1999) suggesting a longer breeding

Common Koel Eudynamys scolopacea 377 season. In Java they breed nearly all year with most egg records in August (Hoogerwerf 1949); in Flores they breed in November (Ottow and Verheijen 1969); in Sumba fledged young are seen in September (Linsley et al. 1998); in Flores they lay in November and December (Verheijen 1964); in Luzon, Philippine Islands, a female had an enlarged ovary in April (Goodman and Gonzales 1990, FMNH). In Sulawesi near Minahassa, they breed in February and March (Stresemann 1940). In Australia, Koels breed from October to February in Queensland, November to February in the Kimberley Division of Western Australia, and December to February in Northern Territory (Vernon 1968, Storr 1977, 1980, 1984b, Maller and Jones 2001). Brood-parasitic. Hosts in Asia are mainly crows. In Pakistan they parasitize House Crows Corvus splendens (Roberts 1991). In India the hosts are House Crows and Jungle Crows (Large-billed Crows) C. macrorhynchos. Koel eggs rarely appear in other species’ nests (Dewar 1908, Lamba 1963, 1976,Ali and Ripley 1969); one Koel was reared by Black Drongo Drongo macrocercus (Smith 1950). In Hong Kong they use Blue Magpie Urocissa erythrorhyncha (Lewthwaite 1995), and in the Bangkok region of Thailand where they parasitized Jungle Crows in the 1920s (Herbert 1924, Aagaard 1930) by 2000 they parasitized two common starlings instead, Common Mynah Acridotheres tristis and White-vented Mynah A. grandis (P. Round). In the Malay Peninsula and Singapore they lay in nests of House Crows, Jungle Crows and mynahs (Medway and Wells 1976, Davison and Fook 1995, Wells 1999; in Singapore mainly House Crows, Wang Luan Keng pers. comm.); in Java they parasitize Jungle Crows and Slender-billed Crow C. enca (Hellebrekers and Hoogerwerf 1967); on Flores they parasitize Jungle Crows (which are also parasitized by Channel-billed Cuckoo Scythrops novaehollandiae) and Flores Crows (Corvus florensis) (Ottow and Verheijen 1969); on Palawan they parasitize mynahs Gracula religiosa (Gonzales 1983). No crows occur on Tinjil and Deli islands, West Java, where Koels may parasitize Black-naped Orioles Oriolus chinensis (Holmes and van Balen 1996). Eggs are gray-blue with small brown or black spots, like crow eggs but smaller, in India 31 ⫻

23 mm (Baker 1934, Lamba 1963), in Java 34 ⫻ 26 mm (Hoogerwerf 1949), in Flores 34 ⫻ 25 mm (Ottow and Verheijen 1969). In India, the female usually lays just after the host has laid its first egg. More than one egg is often laid in a nest, usually 2–3 but also up to 11. The large sets vary in color and may be laid by more than one female (Abdulali 1931, Lamba 1969). The Koel may remove a host egg (Baker 1934). Its shorter incubation (13–14 days, vs 16–20 days in crows) allows the Koel to hatch first, and Koel nestlings lower the breeding success of the host.The Koel nestling is 7 g at hatching and grows to 120–125 g by fledging. The nestling may evict host eggs and young but in other nests the host young may grow up together with the young cuckoo, and two or more cuckoos sometimes fledge from a nest. The nestling period is 19–28 days, and the fledged young cuckoo is fed by its foster parents for another 2–3 weeks (Baker 1934, Lamba 1963, 1969, 1976, Roberts 1991). Begging calls of Koel fledglings are similar to those of the host young (Lamba 1969). Adult female Koels once were said to feed fledged cuckoos (Hume 1873), but this was not seen by Jerdon (1862), Baker (1934), Lamba (1969) or Roberts (1991). Male Koels have not been seen to feed the young. Frequency of parasitism: Koels parasitize 5% of Corvus splendens nests and 0.5% of C. macrorhynchos nests (Lamba 1976). In New Britain, Koels parasitize New Britain Friarbirds (Philemon cockerelli ) (Stresemann 1927a, Meyer 1933). They are mobbed by Yellow-faced Mynahs (Mino dumontii), suggesting a host-parasite relationship (Meyer 1936). The Koel eggs are pink (Meyer 1936). On Sulawesi they parasitize mynahs (Sulawesi Mynah Basilornis celebensis fed a fledgling, and Whitenecked Mynah Streptocitta albicollis is a suspected host) (Stresemann 1940, Andrew and Holmes 1990, Coates and Bishop 1997).The eggs are pink, spotted, 39 ⫻ 24 mm (Schönwetter 1964). In Australia, the hosts are large honeyeaters (Noisy Friarbirds Philemon carunculatus, other friarbird Philemon species and Red Wattlebirds Anthochaera carunculata), Magpie-larks Grallina cyanoleuca, orioles Oriolus sagittatus and Figbirds Sphecotheres viridis which all rear the young Koels, and cuckoo eggs are laid in nests of miners Manorhina spp. and Blue-faced

378 Long-tailed Cuckoo Urodynamis taitensis Honeyeater Entomyzon cyanotis (MacGillivray 1914, Gosper 1962, 1997, Vernon 1968, Goodwin 1974, Storr 1980, 1984b, Crouther 1985, Brooker and Brooker 1989a, Johnstone and Storr 1998, Higgins 1999). Eggs are salmon pink, marbled reddish and violet gray at the large end, 34 ⫻ 24 mm, one egg in a host nest (in contrast to Indian Koel). The female usually removes a host egg. Incubation period ranges from 14–15 days (Gosper 1964, 1997) to 16–17 days (Crouther and Crouther 1984). Nestling Koel evicts the eggs and nestlings of the host with an evicting posture like that of Cuculus and Chrysococcyx, at 48–72 hours after the cuckoo hatches, and does not evict after that period. Nestling Koels are larger than the host nestlings, and they evict their nestmates or outcompete the host nestlings which starve and die, in contrast to Indian Koels which use larger hosts and grow up together in the mixed brood. Nestling growth is rapid; by day 5 the skin is tan, feather tracts emerge and the eyes begin to open, by day 7 the eyes are open, pin feathers brush out on the brow, wings and tail, by day 10 the nestling is well covered above

with feathers. Weight of a nestling Koel in nest of Figbird was 52.5 g at 19–20 days, it stayed in the nest another week and grew more before it fledged (Crouther 1985). The nestling period is 19–21 days in nests of Magpie-larks and 28 days in nests of Figbirds, the difference lying perhaps in the nestling diet, spiders and insects by Magpie-larks and fruit by Figbirds. Nestlings beg by calling, stretching the neck and quivering the wings. Fledglings scramble among branches and remain near the nest and are fed by foster parents for 3–4 weeks after leaving the nest, and other adult birds are attracted to the begging fledged young and feed them.The figbird koel returns to the nest at night for a few days after it first fledges. Fledged young begin to feed on fruit while they are still fed by the foster parents. Breeding success, in observations of 22 eggs, 11 (50%) hatched and 8 (36%) fledged. Of nests that failed, most were either deserted or the Koel egg disappeared (Gosper 1962, 1964, 1997, Beruldsen 1980, Crouther and Crouther 1984, Crouther 1985, Brooker and Brooker 1989a, Higgins 1999).

Genus Urodynamis Salvadori, 1880 Urodynamis, Salvadori, 1880, Orn. Papua e delle Molucca, 1, p. 370.The name Urodynamis refers to the long tail (Gr. oura) and to the similarity in appearance to another large cuckoo, Eudynamys.Type, by original designation and monotypy, Cuculus taitensis Sparrman 1787.The distinctive, long-tailed bird has sometimes been included in the genus Eudynamys (e.g. Mayr

1944a, Mason 1997), but molecular genetics indicates a more recent common ancestor with Channel-billed Cuckoo Scythrops novaehollandiae. Nostril an oval slit, tail wedge-shaped, sexes alike in plumage. Its calls suggest Pachycoccyx. Because it is not a koel, the common name in New Zealand, Long-tailed Cuckoo, is more appropriate than Long-tailed Koel. One species.

Long-tailed Cuckoo Urodynamis taitensis (Sparrman, 1787) Cuculus taitensis Sparrman, 1787, Museum Carlsonianum, 2, p. 32. [Tahiti] Other names: Long-tailed Koel, Pacific Longtailed Cuckoo. Other scientific names: Eudynamys taitensis. Monotypic.

Description

ish streaked buff to rufous, white streak above eye, bold dark brown streak from below eye and down side of neck; underparts white to light rufous with black streaks on side of throat, breast and belly with coarse blackish streaks grading to chevrons on sides of under tail coverts; eye-ring greenish, iris light brown to yellow, bill yellow horn, nostril slit-like, feet olive-green.

ADULT: Sexes similar, above back, wing and long tail brown barred rufous, tip of tail white, small white spots on back and wing coverts, head black-

JUVENILE: Above brown with white and buff spots, head blackish with paler streaking than adult, tail

Long-tailed Cuckoo Urodynamis taitensis 379 above dark brown barred buff to white, underparts buff with black streaks, underparts more rufous than in the adult; eye-ring gray-brown, iris brown, bill brown above and paler horn below, feet horn with yellow edges of scales. NESTLING: Naked, blind at hatching, skin gray-black or dark brown; gape orange or yellow (Wilkinson 1947, Higgins 1999). SOURCES: AMNH, BMNH, CM, FMNH, MVZ, ROM, SMF, UMMZ, USNM, UWBM.

Measurements and weights Wing, M (n ⫽ 23) 183–197 (189.0 ⫾ 3.49), F (n ⫽ 25) 176–196 (185.5 ⫾ 5.1); tail, M 211–236 (226.8 ⫾ 6.0), F 204–237 (220.2 ⫾ 7.6); bill, M 29.5–35.0 (33.4 ⫾ 1.3), F 32.0–35.7 (33.5 ⫾ 0.7); tarsus, M 33.4–37.3 (34.9 ⫾ 1.1), F 32.5–36.6 (34.7 ⫾ 1.0) (Higgins 1999). Weight, M (n ⫽ 9) 74–153 (110.4), F (n ⫽ 10) 93–148 (124.2) (Higgins 1999); wintering M “extremely fat” 151 (UWBM). Wing formula, P8 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 10 ⬎ 1.

Field characters Overall length 38–42 cm. A long-tailed cuckoo with rufous markings, spotted and white and rufous on the head, back and streaked black below, like a young hawk. with rapidly beating, short, pointed wings.

brown barred wings, It flies

Voice Whistled “ouit-ouit”, a magpie-like chatter “wheet-wheet-wheet” and shrieks, a loud grating whistle “whiz-z-z-z-z-t” or “zzwheesht”, shrill and upslurred (2 to 5 kHz). Chatter is given in flight and on a perch. Other calls are twittering from premigratory flocks, an alarm whistle, and a subdued mutter (Higgins 1999). Begging calls of young reared by Whiteheads Mohoua albicilla are similar to calls of Whitehead, “whir-r-r-r”, like winding a mechanical watch. Fledglings trill, like fledged young host Whiteheads and Yellowheads Mohoua novaeseelandiae (Wilkinson 1947, McLean and Waas 1987, Buckingham and Jackson 1990).

Range and status Breeds in New Zealand (North and South Islands), Great Barrier I, Little Barrier and Stuart Islands (Falla et al. 1978, McLean 1988, Higgins 1999). Migratory, they occur during the austral winter on islands in the Pacific in Oceania east to the Pitcairn Islands, Marquesa Islands, Society, Samoa and Cook Islands, Tonga, Fiji (Clunie and Morse 1984), Solomon Islands (on small outlying islands, Coates 1985), Vanuatu, Norfolk I, the Bismarck Archipelago (Meyer 1936), New Guinea (on small outlying islands in the NE and SE, where it is rare; it is not known on the New Guinea mainland); and north of the equator on Palau, the Caroline Islands (Yap, Lukunor, Truk, Ponapé, Kusaie) and Marshall Islands ( Jaluit, Elmore, Auru, Wotze, Bikini) (Baker 1951, Ellis et al. 1990, Buden 1999, Higgins 1999). Adults arrive in New Zealand in October and depart in January and February, migrate through Lord Howe and Norfolk Islands, and New Caledonia, and winter in Melanesia, Micronesia and Polynesia east of Fiji, where they remain from May to September for the austral winter. In the winter range they show up as early as February in the Caroline Islands (Baker 1951). On Tahiti, Pitcairn and Henderson Islands they occur from June to September (Bogert 1937, Thibault and Rives 1975, Brooke 1995). In New Zealand a few stay behind on North Island in winter. Vagrant on Chatham Islands and Kermadec Islands; perhaps a breeder on Norfolk I (Schodde et al. 1983, Higgins 1999). Numbers have decreased with the loss of forests and populations of songbird hosts in New Zealand. Long-distance migrants, they move at night and call in flight.

380 Genus Scythrops

Habitat and general habits

Breeding

Forest canopy, native forests, and introduced pines, overgrown coconut plantations and gardens, and lowland scrub into hill forests. Breeding distribution of cuckoos in New Zealand coincides with that of the host species, mainly in forested mountains (Cometti 1986, Heather and Robertson 1997, Higgins 1999). On islands in migration and in the winter range they occur in lower vegetation, including scrub. Feed singly in trees and shrubs, actively glean large insects from twigs, branches and bark, dashing from one tree to another; in the open they fly rapidly and directly with flapping wingbeats (Holyoak 1980). From their hawk-like appearance they are known as “sparrowhawk” (Williams 1960).They call and feed at night as well as in daytime, but on Little Barrier I they feed only at night (Higgins 1999). Occur singly and in groups of 2–5.

Season, is November–December or early January; hosts that nest earlier in the spring season escape the cuckoo. Brood-parasitic. Hosts are several small songbirds including Whitehead Mohoua albicilla on North I and Little Barrier I, and Yellowhead Mohoua novaeseelandiae and Brown Creeper Finschia novaeseelandiae on South I, and occasionally other species including Silvereyes Zosterops lateralis and introduced House Sparrow Passer domesticus (Falla et al. 1978, McLean 1982, 1988, Cometti 1986, Higgins 1999). Nests are parasitized over a wide range of heights, from less than 1 m to 35 m in a tree cavity. Eggs are whitish, variably marked with purplish brown and gray, 23.5 ⫻ 17 mm (Oliver 1955, McLean 1988; AMNH). Incubation period is about 16 days. Nestlings at times evict the eggs of the host. At other times, mixed-species broods of cuckoo and Yellowhead fledge both the cuckoo and the host young. The nestling period is 21 days in the nest of the Whitehead (longer than the 17 days of the nestling Whitehead) to 29 days in the nest of the Yellowhead.Young cuckoos are fed for at least two weeks after fledging. The fledglings beg from birds and species other than their own foster parents and are sometimes fed by these birds. Whiteheads sometimes feed two young cuckoos, probably because the second was adopted, as Whiteheads with fledglings often form mixedfamily flocks. Breeding success, 6 (6%) of 95 Yellowhead nests were parasitized in one site, 2 (8%) of 24 in another site. On Little Barrier I, 16% of Whiteheads flocks that were seen with fledged young had a fledged cuckoo (McLean 1982, 1988, McLean and Waas 1987, Higgins 1999).

Food Mainly large invertebrates such as insects and crabs; skinks, geckos, eggs and nestling birds; a predator on nestlings even when not in breeding. Also takes berries and fruit (Oliver 1955, Dow 1972, Gill 1980b, McLean et al. 1986, Higgins 1999).A nestling reared by Whiteheads was fed cicadas and caterpillars (Wilkinson 1947).

Displays and breeding behavior Males sometimes display in groups that last a few days, then break up and later re-form in other places (perhaps the same males, no observations have been made on marked individual birds). A male perches for long periods giving screech and chatter, in an upright posture with tail fanned, wings extended upward, bent and slowly flapped (McLean 1988).

Genus Scythrops Latham, 1790 Scythrops Latham, 1790, Index Ornithologicus, 1, p. 141. Huge cuckoo with gray plumage, nostril oval. Type, by monotypy, Scythrops novaehollandiae

Latham. The name (Gr. skuthros, angry, sullen, scowling; ops, the face), describes the huge curved bill and grim appearance. One species.

Channel-billed Cuckoo Scythrops novaehollandiae 381

Channel-billed Cuckoo Scythrops novaehollandiae (Latham, 1790) Scythrops novae Hollandiae Latham 1790, Index Ornithologicus, 1, p. 141. (New Holland ⫽ New South Wales) Polytypic. Three subspecies. Scythrops novaehollandiae novaehollandiae Latham 1790, Scythrops novaehollandiae fordi Mason 1996, Scythrops novaehollandiae schoddei Mason 1996.

Description ADULT: Sexes similar, male above head light gray, back and wing coverts darker gray with black edge on tip of feathers forming bars or scallops, primaries and secondaries gray with blackish tips, tail graduated, the feathers gray with broad subterminal black band and white tip, T1 otherwise unbarred, T2 to T5 inner vane whitish with thin black and buff bars, outer vane and base of inner vane gray, face gray; underparts, breast to belly and under tail coverts light gray, flanks barred darker gray, underwing coverts light gray, wing below with black trailing edge on middle third of primaries, female with underparts paler than male, more extensively barred on flanks and belly; in both sexes bare skin around eye reddish, extends through lores to nostril, iris red, bill huge, two broad grooves or channels on each side of upper mandible extend from base of bill to near tip, pale gray with yellow tip (distal 20–30% in male, 30–50% in female) and a bright pink tip in spring, bill slightly grooved, feet gray. JUVENILE: Head and neck yellowish buff, large buff spots on wing, the primaries and secondaries subterminally black and the tip whitish, tail averages longer than adult, underparts more strongly barred than in adult, underwing below with pale trailing edge; bare skin around eye grayish brown, eye-ring darker and narrower than in adult, iris brown, bill short and narrow and lacks grooves, dark gray to pinkish, feet pale gray green. NESTLING: Naked at hatching, skin bronze; mouth pinkish red, bill dark blackish brown, paler near the tip.

SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MSNG, MZB, RMNH, ROM, SMTD, UMMZ, ZMB, ZMUC.

Subspecies and history Scythrops novaehollandiae novaehollandiae Latham 1790; bill smaller; Australia; Scythrops novaehollandiae fordi Mason 1996; bill larger; Sulawesi, Buton I,Tukangbesi Is, Banggai Is, Sula Is; Scythrops novaehollandiae schoddei Mason 1996; bill larger, plumage of back more strongly scalloped; Bismarck Archipelago, northern Melanesia. Channel-billed Cuckoos were collected by the first European settlers at Sydney, New South Wales, and were illustrated in two publications (White’s Journal, Phillips’s Voyage) where called “Anomalous Hornbill” and “Psittaceous Hornbill”. Descriptions by Latham in 1790 gave the common name “Channel Bill” and noted their seasonal appearance. Latham noted their diet of eucalypt seeds (red-gum and peppermint), the strong horny bill to crush the capsules, and the short tongue to pick the seeds from the capsules. He described the cuckoos as nesting and rearing their own young. John Gould first reported that young Channel-billed Cuckoos were reared by other species (Gould 1845b, Mathews 1918). Two subspecies were described by Mason (in Mason and Forrester 1996), S. n. fordi in Sulawesi and the Moluccas, and S. n. schoddei in northern Melanesia. S. n. fordi was said to be larger than Australian birds especially the bill, and S. n. schoddei to be larger than Australian birds and to have the plumage more scalloped above. No consistent differences in wing or bill size or plumage pattern are apparent in specimens in all museums, but these characters vary with sex and some specimens may be incorrectly sexed. In the descriptions of these subspecies, no known breeding birds in Sulawesi, Flores or Melanesia were examined, and birds that did not fit the overall pattern were excluded from the results (Mason and Forrester 1996). Consistent with the descriptions, a male from the northern

382 Channel-billed Cuckoo Scythrops novaehollandiae Melanesian island of Djaul (ZMUC 1.3.1982.567, Noona Expedition, 9 June 1962) was large (wing 357, tail 322, bill 87, tarsus 40, testes 6 mm). The measurements however overlap between birds of Australia and birds of Wallacea and northern Melanesia (Mason and Forrester 1996, Higgins 1999). Birds of the eastern Lesser Sunda Is may include migrants from Australia as well as local breeders.The only adult specimen seen from Flores (RMNH #C.35) taken in 1890 is of unknown month and sex and is indeterminate in size (wing 342, bill from nostril 76, bill from head 85) and of uncertain subspecies identification; the only juvenile (ZMB 17670) is not fully grown.

Measurements and weights S. n. novaehollandiae, Australia: Wing, M (n ⫽ 21) 337–367 (353.9 ⫾ 7.8), F (n ⫽ 9) 336–368 (346.8 ⫾ 9.1); tail, M 253–278 (266.5 ⫾ 7.4), F 253–276 (262.8 ⫾ 7.2); bill, M 83.7–93.5 (86.4 ⫾ 2.3), F 70.3–79.0 (74.5 ⫾ 3.0); tarsus, M 41.8–49.8 (45.8 ⫾ 2.3), F 38.7–46.3 (44.0 ⫾ 2.0) (Higgins 1999); Moluccas and New Guinea, wing, M (n ⫽ 2) 349–358 (358.5), F (n ⫽ 4) 340–352 (346); tail, M 300, F 280–304 (285); bill from nostril, M 65–72 (68.5), F 60–64 (61.5), bill from head, M 72–80 (76), F 68–75 (72.3); bill depth (from ridge on upper mandible anterior to the nostril, to base of feathered chin), M 28, F 26–29 (27.8); tarsus, M 46, F40–48 (44) (MZB); S. n. fordi, Sulawesi: Wing, M (n ⫽ 17) (361.5 ⫾ 4.3), F (n ⫽ 7) (347.8 ⫾ 5.0); tail, M (274.0 ⫾ 7.4), F (265.5 ⫾ 9.0); bill from nostril, M 71.7 ⫾ 3.9), F (59.5 ⫾ 2.1) (Mason and Forrester 1996); Sulawesi and Sula Is, M (n ⫽ 3) 350–365 (356.3), F (n ⫽ 4) 344–352 (350.8); tail, M 260–300 (279.3), F 262– 300 (284.5); bill from nostril, M 65–76 (72.3), F 67–73 (69.3), bill from head, M 75–87 (83), F 78– 87 (82.5); bill depth M 29–34 (32.3), F 43–48 (45.7); tarsus, M 42–50 (47.3), F 43–48 (45.7) (MZB); Northern Melanesia: Wing, M (n ⫽ 4) (352.5 ⫾ 8.5), F (n ⫽ 3) (346.5 ⫾ 2.6); tail, M (275.0 ⫾ 7.1), F (274.5 ⫾ 3.5); bill from nostril, M (80.5 ⫾ 3.3), F (79.7 ⫾ 6.6) (Mason and Forrester 1996). Weight, Australia: M (n ⫽ 12) 560–935 (708.2 ⫾ 113.5), F (n ⫽ 12) 560–800 (660.0 ⫾ 83.7); before post-breeding migration, M, October–November (n ⫽ 5) (665.8 ⫾ 50.9); January–February (n ⫽ 6)

(746.7 ⫾ 67.5), March (n ⫽ 1) 935.1 (Higgins 1999); Halmahera: M (n ⫽ 2) 342, F (n ⫽ 2) 346– 370 (358) (Ripley 1964), Ambon, F (n ⫽ 1) 620 (MZB). Wing formula, P8 ⬎ 9 ⱖ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 10 ⬎ 2 ⬎ 1; P8 ⬎ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 60 cm. Huge, deep, channeled and decurved bill, gray plumage, long pointed wings, and a long slender tail with bands on the tip. First-year adults often have a mixed plumage of old juvenile feathers and new adult feathers. Looks like a large gray hawk or a hornbill, but the tail is graduated in shape and the bill is much larger than that of a hawk, and the bird often is in a fruiting tree. Flight is slow and direct with slow and regular wing beat, high above ground or water, the flight silhouette a “southern cross” or “Maltese cross”. Channel-billed Cuckoos are the largest brood-parasitic bird in the world.

Voice (1) a loud, drawn-out, rising guttural, nasal scream, “crrRRRROW” or “crrRRRRAY”, often repeated rapidly in an ascending series, and (2) a descending series of shorter disyllabic downslurred screams, “CLOo-CLOo-CLu-Clu-clu- . . .”, like a bubbling trumpet, sometimes two or more birds call wildly at the same time with both calls, and (3) loud raucous, repeated coarse notes “gaak” all on the same pitch, at 2 notes in 3 sec. Female gives a reedy trumpeting when male approaches with food in courtship feeding, and both birds give guttural calls (Hindwood and McGill 1951, Johnson 1983, Higgins 1999). Noisy in early morning, late afternoon and evening and into night. Sometimes two or more birds call together giving both the rising and downslurred calls (Buckingham and Jackson 1990, Coates 2001). Birds call when perched and in flight, and they are usually quiet when feeding (Stresemann 1940, Slater 1971, Beehler 1986, Coates and Bishop 1997, Pizzey and Knight 1998, Higgins 1999). At roost, birds sound like a herd of pigs squealing, fighting and grunting; in morning they croak hoarsely when leaving the tree (LeCroy and Peckover 1983). Calls in Sulawesi (Coomans de Ruiter 1950) are similar to those in Australia. Begging calls of the young fledged cuckoos are loud, screaming and shrieking, said to be similar to begging

Channel-billed Cuckoo Scythrops novaehollandiae 383 calls of their hosts Pied Currawong Strepera graculina, Australian Magpie Gymnorhina tibicen and crows, but louder and more raucous (Goddard and Marchant 1983); another account noted that young cuckoos reared by crows give falsetto “Punch-and-Judy” noises unlike the “aark aark” of young crows (Salter 1978). Adult-like cuckoo calls were given by a captive-reared bird a month after it fledged (Higgins 1999).

Range and status Australia, Lesser Sunda Is, Sulawesi and nearby islands, Halmahera and the Moluccas, New Guinea and nearby islands, and the Bismarck Archipelago. In Australia, they occur seasonally in Northern Territory from October to April, in Queensland from September to April, and in some years they occur into the heart of Australia near Alice Springs and breed there. Cuckoos are migratory and seasonal appearing first on the north coast.They are diurnal migrants in flocks of up to 30 birds. Passage movements are observed in the Torres Strait during August and September; migratory movements near Darwin and Arnhem Land in Northern Territory also indicate movement between Australia and New Guinea. Flocks fly northward in coastal NE Queensland in March. Nearly all records in Northern Territory are between October and April (Bravery 1970, Storr 1977, 1984b, Mason and Forrester 1996, Nielsen 1996, Higgins 1999). A sight record from Ashmore Reef in October suggests migration across the Banda Sea between Western Australia and the Moluccas.There is a long-distance recovery of a banded bird, marked in Ourimbah,

NSW and found at Mollymook, NSW, a distance of 235 km. A few birds remain in the breeding range in the dry season along rivers in the Kimberley Division of Western Australia, the Darwin area and Kakadu NP in the Northern Territories, and Torres Strait and northern Queensland (Higgins 1999). In New Guinea they occur in all months with most records between February–March and October–November during the Australian nonbreeding season.They may breed; there are no breeding records (Coates 1985, 2001). Birds range to the northwest and north islands of New Guinea from Misool, Waigeo, Kofiau, Manam and Karkar. In the Bismarck Archipelago the birds are seen mainly from February to May (Coates 1985). Nevertheless, on New Britain channel-bills have been seen and collected in all seasons, in February, April, May, June, October and December (LeCroy and Peckover 1983, AMNH), a chick and eggs were taken from the nests of a crow (Meyer 1927, 1933). In the Louisade Archipelago they appear in diurnal migration in October (LeCroy and Peckover 1998). In the Solomon Is there are observations on Savo, Simbo and Rennell, all of these islands isolated and species-poor, with no suggestion of breeding (Diamond 2002). In Wallacea, in northern Sulawesi Channel-billed Cuckoos are noisy and conspicuous, occurring in all months (Mason and Forrester 1996); the northern limits are Siau and Talaud islands (Riley 1997). In the Moluccas most records are from April to September and breeding has not been reported. In Halmahera, Ripley (1959) saw small groups in September flying north, not south towards Australia, perhaps moving between feeding and roosting sites. In the Lesser Sundas they breed in Flores and perhaps east to Tanimbar, in Sumba they are wet-season breeders or winter visitors (Linsley et al. 1998). Rare in Timor,Wetar and Alar (perhaps migrants from Australia, Monk et al. 1997).Vagrant in Tasmania, New Zealand, Norfolk I and New Caledonia (White and Bruce 1986, Bruce 1987, Noske 1994, Mason and Forrester 1996, Coates and Bishop 1997, Heather and Robertson 1997,Verhoeye and Holmes 1998, Higgins 1999). Channel-bills are uncommon in Australia, are locally common in austral winter in New Guinea, occur all year in New Britain, are uncommon in most of Wallacea and are locally common in Sulawesi and Flores.

384 Channel-billed Cuckoo Scythrops novaehollandiae

Habitat and general habits Forests and woodlands where figs Ficus spp. occur, canopy trees at edge of forest and along river, mainly Eucalyptus spp., other open woodlands with tall trees, figs, secondary forest, mangroves. Occur mainly in lowlands in subhumid and semi-arid regions, often near the coast, and on coastal islands, scarce and patchy in arid zone of Australia where they occur mainly in waterside vegetation (Storr 1980, 1984b); in lowlands and hills to 1000 m in Sulawesi and to 340 m in Flores. They feed in groups of up to 20 birds in the canopy of fruiting trees (Coates 1985), rarely in mixed-species flocks with other fruit-eating birds (Higgins 1999). Shy, though noisy and conspicuous with raucous calls. Flight is strong and direct, with strong wingbeats 1/sec and speed 75 km/h at 30–80 m above the ground or sea. They are active at night, calling, flying about and feeding on fruit (Mathews 1918, Higgins 1999). Local movements occur between feeding and roosting areas, as in northwest Torres Strait where they feed on islands and roost on the New Guinea mainland (Draffan et al. 1983). Occur in singles, pairs and small parties and common in flocks in the breeding season.They gorge on figs in the weeks before they migrate (Crouther 1980).

Food Fruit, mainly figs, mistletoe berries, mulberries Morus nigra, Eucalyptus seeds, and large insects including orthoptera (stick insects, locusts), caterpillars and beetles, and the eggs and young of birds (Gould 1845b, Dahl 1899, MacGillivray 1914, Mathews 1918, Hindwood and McGill 1951, Lord 1956, Gilliard and LeCroy 1967a, Goodwin 1974, Salter 1978, Blakers et al. 1984, Kloot 1993, Larkins 1994b, Johnstone and Storr 1998, Higgins 1999). Nestlings reared by currawongs are reared on fruit, nestlings reared by crows are fed insects (caterpillars, beetles and grasshoppers) and carrion.

Displays and breeding behavior Occur singly, in pairs and in flocks of up to 10 birds (Stresemann 1940, Goodwin 1974, Higgins 1999). In aggressive behavior, a cuckoo stands high on extended legs with the body crouched forward, hops toward another channel-bill, croaking and with wings partly extended and held low, then it

supplants and chases the other bird and it returns to tree (LeCroy and Peckover 1983). In courtship feeding, the breeding male presents a large insect to the female, she crouches and spreads her wings, the male then mounts and mates with her, and the female takes food from the male before they mate or during copulation (Hindwood and McGill 1951, Goddard and Marchant 1983, Johnson 1983). Cuckoos approach the nesting host in pairs, the male distracts the host while the female lays in the unattended nest (Crouther 1980, Larkins 1994a,b).

Breeding They sing and breed in N Sulawesi (Dumoga Valley) from late February to June with fledglings in March, in north-central Sulawesi they breed in November (Koelawi), at Lore Lindu in January, on Boni I a nestling was nearly grown with wing 330 mm in March (Mason and Forrester 1996, RMNH) and on Butung I they call early in the rainy season (October and November) and gonads are active in September (van Bemmel and Voous 1951). In the Sula Islands they breed in December (Mangole, nestling, RMNH). In Flores they begin to call in October and breed from November to March (Verheijen 1964, Ottow and Verheijen 1969). In New Britain, a laying female was taken in December (AMNH 334029) and males had large testes in December and February (AMNH 334028, 777854); although O. Meyer did not note a date for his cuckoo egg, he found the host crows bred from November to April (Meyer 1933, 1936, LeCroy and Peckover 1983). In Australia they breed during the austral spring and summer, in the Kimberley Division of Western Australia from November to January, in Queensland from October to January, at Alice Springs in January and in southeastern Australia from October to December (Storr 1980, 1984b, Strahan 1994, Verhoeye and Holmes 1998, Higgins 1999). Adults are often mobbed and chased by crows (Stresemann 1940, Salter 1978, Beehler et al. 1986, Coates and Bishop 1997). Brood-parasitic. The hosts are large songbirds, half the size to same size as the cuckoos, the host size ranges 387–615 g (Storr 1984, Brooker and Brooker 1989a). In Sulawesi they parasitize Slenderbilled Crow Corvus enca (Meyer 1879,Watling 1983, Rozendaal and Dekker 1989, Mason and Forrester 1996), in Flores and Sumba the Flores Crow

Asian Emerald Cuckoo Chrysococcyx maculatus 385 C. florensis and Jungle Crow C. macrorhynchos (Ottow and Verheijen 1969, Coates and Bishop 1997). In New Britain they parasitize Torresian Crow C. orru (Meyer 1933, Schönwetter 1964). In Australia the hosts include Torresian Crow Corvus orru and other crows, Pied Currawong Strepera graculina, butcherbirds Cracticus spp., Australian Magpie Gymnorhina tibicen and Magpie Lark Grallina cyanoleuca; most records are currawong and crow (Goddard and Marchant 1983, Higgins 1999). Eggs in Sulawesi are undescribed; eggs in Flores are light blue-green with violet-gray flecks like eggs of Jungle Crow host (Ottow and Verheijen 1969); eggs in New Britain are muddy yellow with green and gray-brown flecks, 40 ⫻ 29 mm, the same size as crow eggs which are greenish with darker flecks (Meyer 1927, 1933), and 40 ⫻ 27 mm, dull pinkishgray with violet-gray spots (Schönwetter 1964); eggs in Australia are dull white, blotched reddish or brown, like currawong eggs, 41.7 ⫻ 28.9 mm (Schönwetter 1964, Goddard and Marchant 1983). The adult female cuckoo often removes, cracks or

dents a host egg when she lays her own, and she sometimes eats a host egg (Goddard and Marchant 1983, Larkins 1994b). Female lays 1–5 (often 2) eggs in a nest, 52% of parasitized nests have 2–5 cuckoo eggs, laid by one female or by more than one, to judge from the variation in color and pattern of the eggshells (Storr 1984b, Brooker and Brooker 1989a). The incubation period is unknown. Nestling hosts usually disappear within a week.The nestling cuckoo may crowd and starve them without evicting them, as more than one cuckoo (up to 5) may fledge from a single nest, and occasionally young crows fledge as well. The cuckoo fledges in 17–24 days and begs from its foster parents, its body horizontal, wings quivering and held out from the sides. The young cuckoo depends on its foster parents for another month, and a fledgling taken from its crow foster parents stayed with its human foster parent for 30 days, flying free and returning to be fed (Meyer 1879, Berney 1907, MacGillivray 1914, Lord 1956, Frauca 1967, Storr 1980, Goddard and Marchant 1983, Higgins 1999).

Genus Chrysococcyx Boie, 1826 Chrysococcyx Boie, 1826, Isis von Oken, Band 2, col. 977. Small brood-parasitic cuckoos with whistled songs, often with glossy plumage of bronze or green, and with round nostrils.The genus name Chrysococcyx (Gr. khrusos, gold; kokkux, the cuckoo) refers to the metallic sheen of the plumage. Type, by monotypy, Cuculus cupreus Latham ⫽ Cuculus cupreus Shaw. Australasian species are often separated as a genus Chalcites Lesson 1830 (type species, C. lucidus plagosus). Nevertheless, molecular genetics indicates that Asian Emerald Cuckoo C. maculatus and Violet Cuckoo

C. xanthorhynchus are more closely related to the African glossy cuckoos than to the other Australasian glossy cuckoos, as suggested also by Friedmann (1968). Long-billed Cuckoo C. megarhynchus of New Guinea, formerly listed in a separate genus Rhamphomantis, is in this clade and resembles other Chrysococcyx in its glossy bronze plumage (brighter than in C. osculans, and similar in color to Australian-Papuan C. minutillus poecilurus) and its slurred and trilled whistled songs. No anatomical material of C. megarhynchus is available for comparison. Fourteen species.

Asian Emerald Cuckoo Chrysococcyx maculatus (Gmelin, 1788) Trogon maculatus Gmelin, 1788, Systema Naturae, 1, part 1, p. 404. (Ceylon) Other names: Chalcites maculatus (Gmelin). Monotypic.

Description ADULT: Male, above, head to back and rump glossy green, wing blackish with outer vane of flight feathers glossy green, tail glossy green, inner tail

386 Asian Emerald Cuckoo Chrysococcyx maculatus feathers T1, T2 unbarred and dark to tip, T3 to T5 with white tip, T5 outer tail feathers with incomplete narrow white bars; underparts, chin to upper breast glossy green, belly white with black to green and bronze bars, under tail coverts green with white spots, underside of wing has a white band across the inner primaries and secondaries, under wing coverts white with black bars; eye-ring orange red, iris red to red-brown, bill orange with black tip, feet green. Female, crown to upper back rufous or rufous lightly barred bronze, back and rump unbarred bronze-green, wing coverts bronze-green, wing blackish with outer vane of flight feathers bronze-green, tail green, rufous and white, central feathers T1 and T2 bronze-green, outer feathers T3 to T5 barred green and rufous with black subterminal band, T4 with white tip and T5 with white spots and tips; underparts, white barred with bronze, primaries and secondaries show rufous window on trailing vane in flight from below with a narrow white band anterior to rufous window, under wing coverts white with black bars; iris brown, bill yellow, black at tip, feet brownish green. JUVENILE: Head rufous or rufous barred white or barred green and white, or head and upper back unbarred rufous, back and wing coverts barred rufous and bronze-green, or rufous and brown, bars indistinct, wing flight feathers blackish with outer vane of flight feathers bronze green, tail green and rufous with white spots,T1 and T2 unbarred green, T3 to T5 barred green and rufous,T5 barred brown and rufous with white spots; underparts, throat and breast rufous barred with dark bronze, belly white barred with brown, underside of wing as in adult female, under tail coverts white barred bronzeblack; bill black, lower base flesh, feet gray. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, FMNH, MCZ, ROM, UMMZ, USNM, SMF,YPM, ZMUC.

15.9 (14.9 ⫾ 0.5), F 14.3–16.4 (15.0 ⫾ 0.8); tarsus, M 13.8–16.2 (15.0 ⫾ 0.9), F 14.0–16.0 (14.8 ⫾ 0.6) (AMNH, ANSP, MCZ, ROM, UMMZ, YPM). Weight, M (n ⫽ 3) 23–27 (25.0), F (n ⫽ 1) 30 (BMNH). Wing formula, P8 ⱖ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 10 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 17 cm. Small cuckoo with glossy green plumage, barred underparts and a red bill; female has rufous head and nape (coppery to graybrown in C. xanthorhynchus) without white speckling on face and a yellow bill, tail similar to C. xanthorhynchus (underparts more finely barred and tail more rufous than C. minutillus), in flight shows a white wing stripe; juveniles have whitish throat (rufous wash in C. xanthorhynchus).

Voice Song, a loud whistle of 3 notes “kee-kee-kee”, given rapidly; a “chweek” given in fight; a trill or high rattle of 5–6 notes, and a loud whistled twitter (Betts 1947, Ali and Ripley 1969, King and Dickinson 1975, Robson 2000a). Sings at night (Stevens 1925).

History Named by Gmelin as a spotted cuckoo (“Trogon maculatus”) from the illustration of a cuckoo with spotted wing coverts, “le curucui tacheté” (“Spotted Curucui”), drawn by Peter Brown (1776, p. 28, pl. 13, fig. 2).The illustration was copied from a drawing received from Gideon Loten, who was Resident of Java and Ceylon, and Governor of Ceylon (Brown 1776). Brown’s description (p. 28) reads “Ceylon. From Governor Loten” and the figure is recognizable as a glossy green cuckoo, the feet are zygodactyl, the bill is yellow (as in adult females) (“brown” in Brown 1776), the underparts are “pale brown, barred with dusky,” the wing has large white spots and the tail is barred. There are no other records in Ceylon (Legge 1880, Kotagama and Fernando 1994).

Measurements and weights Wing, M (n ⫽ 21) 104–113 (108.8 ⫾ 2.5), F (n ⫽ 15) 103–114 (109.1 ⫾ 3.7); tail, M 66–76 (70.8 ⫾ 3.1), F 66–78 (70.8 ⫾ 3.3); bill, M 14.0–

Range and status Southern and southeast Asia in northern Pakistan and Indian Himalayas, Nepal, Sikkim, Bhutan,

Asian Emerald Cuckoo Chrysococcyx maculatus 387 the leaves, and take insects on the wing (Stevens 1925).

Food Insects, including ants, caterpillars and bugs (Hume 1888, Stevens 1925).

Breeding

Arunachal Pradesh, Assam, Bangladesh, SE Tibet and southern China (Sichuan, Yunnan, Hunan, Hainan), Burma, Bay of Bengal islands, NW Thailand, Laos, Cambodia and Vietnam. Migratory and seasonal in northern India, Nepal and Tibet (Singh 1994, Grimmett et al. 1999); resident in Yunnan and Hainan (Guangdong et al. 1983, Cheng 1991,Yan 1995,Thomas and Poole 2003). In India, they are rare and not common anywhere (Legge 1880, Hume 1888, Stevens 1925, Ali and Ripley 1969, Grimmett et al. 1999). In winter they are uncommon to rare in N India ( January in Sikkim, ROM; November and March in Lushai Hills, UMMZ; March at Adyar, near Madras, Dique 1880), and in the Nicobar Islands (February at Teressa I, Abdulali 1964, 1967), Thailand, the Malay Peninsula and Sumatra (van Marle and Voous 1988, Lekagul and Round 1991,Wells 1999).

Habitat and general habits Dense evergreen forests, broad-leafed secondary forest, freshwater swamp forest, plantations and gardens. They occur in lowlands to 2500 m in Nepal, 2200 m in Bhutan and Arunachal Pradesh, and 1800 m in Thailand ( Jerdon 1862, Stanford and Ticehurst 1939, Medway and Wells 1976, Inskipp and Inskipp 1985, Smythies 1986, Lekagul and Round 1991, Singh 1994, Grimmett et al. 1999, Robson 2000a). Altitude tends to be lower than that of Violet Cuckoo C. xanthorhynchus; both cuckoos are within 600–1500 m in the plains and hills of Assam (Baker 1927). They feed in outer branches and foliage of the tree canopy, searching through

In India the egg records are from April to July (Baker 1908a, 1934, Stevens 1925, Ali and Ripley 1969). Brood-parasitic. Hosts in India, Yellowbacked Sunbird Aethopyga siparaja and Little Spiderhunter Arachnothera longirostra (Baker 1918). Baker reported that Mr. Primrose reared a young “Emerald Cuckoo Chrysococcyx maculatus” from the nest of the sunbird where he had found cuckoo eggs in an earlier season (Baker 1934); and Chestnut-crowned Warbler Seicercus castaneiceps (Stevens 1925). In the breeding season the adults are seen in pairs (Stevens 1915, 1925). Eggs are whitish, blotched with light brown or reddish brown, like those of spiderhunters, 17 ⫻ 12 mm, narrow at one end (Baker 1906, 1908a, 1942, Harrison 1969, Becking 1981). Baker’s identification of eggs was reviewed and questioned by Harrison (1969) and Becking (1981).The identification was based in part on the occurrence of two kinds of small eggs, thought to be of the C. maculatus and the C. xanthorhynchus; on an egg observed to hatch and the nestling of which developed its plumage; and on the local distribution of cuckoos, as C. maculatus tends to occur at higher elevations (Baker 1934, 1942, Harrison 1969). The identification of young cuckoos as well as of their eggs has long been confused. Baker (1906) reported a nest of Grey-throated Babbler Stachyris nigriceps with two eggs of the babbler and one of a cuckoo, and an adult Emerald Cuckoo caught in a noose at the entrance of the nest; the babbler is otherwise known as a host of the Fork-tailed Drongocuckoo Surniculus dicruroides. The cuckoo egg was much larger than the known eggs of either glossy cuckoo (Baker 1908a). The identification in a Crimson Sunbird Aethopyga siparaja nest, of an egg that hatched and developed into a juvenile cuckoo (Baker 1908a,b) is questioned because the juvenile may have been misidentified and the specimen has not been located (Becking 1981), nor has the

388 Violet Cuckoo Chrysococcyx xanthorhynchus juvenile identified by Stevens (1925) from a nest of Seicercus castaneiceps warbler. Stevens (1924, p. 1026) also suggested that only male cuckoos fledge from the nests of warblers, and only female cuckoos from the nests of sunbirds. In fact, no sex difference in juvenile plumage is known in any species of glossy cuckoo. Stevens based his identification on Baker’s descriptions, and supposed that young with the chestnut head were female (later specimens showed variation with either sex sometimes having a chestnut head). He later noted (Stevens 1925, p. 666) that Baker had transposed the descriptions; Baker (1922, 1927) in fact had described the adult female C. maculatus as rufous barred black below, and the adult female C. xanthorhynchus as white barred

black. In neither species is the female rufous on the underparts; juvenile C. maculatus are rufous barred black on the throat and breast. In addition to the identification of young cuckoos, some egg records of this cuckoo or of C. xanthorhynchus may be eggs of Cacomantis merulinus, whose chicks have been confused with the chicks of glossy cuckoos, as described in the next species. Baker (1927) mentioned Stevens’s record of this cuckoo in the warbler nest, but he did not later include the record (Baker 1934). Incubation and nestling periods are unknown. The nestling cuckoo is thought to evict the host egg, as nests with a cuckoo chick (this species?) have only the cuckoo with no host eggs or young (Baker 1907, 1908a).

Violet Cuckoo Chrysococcyx xanthorhynchus (Horsfield, 1821) Cuculus xanthorhynchus Horsfield, 1821, Transactions of the Linnean Society of London, 13(1), p. 179 ( Java). Polytypic. Two subspecies. Chrysococcyx xanthorhynchus xanthorhynchus (Horsfield, 1821); Chrysococcyx xanthorhynchus amethystinus (Vigors, 1831). Other names: Chalcites xanthorhynchus (Horsfield, 1821)

Description ADULT: Male, head to back and rump glossy violet, wing blackish with outer vane of flight feathers glossy purple, tail blackish glossed violet, inner tail feathers T1 and T2 dark to tip, T3 to T5 tipped white, T5 outer tail feathers T5 dark barred white on outer vane and barred brown on inner vane; underparts, chin to upper breast bright violet, belly white with broad black violet or green bars, under wing coverts white with black bars; eye-ring red, iris red to brown, bill yellow with base orange-red, feet olive green. Female, above greenish bronze, crown coppery to gray-brown, back greenish bronze, face with small white spots around eye and sometimes on forehead, tail greenish to greenish rufous, the inner tail feathers T1 and T2 with an obscure black subterminal band and buff tip, other tail feathers rufous with greenish bars and white spots on outer vane,T4 and T5 tipped white; underparts, chin and throat to breast, belly and under tail coverts whitish

(or rufous wash whitish) with bronze-green bars, wing with flight feathers dark brown, inner vane of primaries rufous, under wing coverts white with black bars; eye-ring red, iris brown, bill dark brown with an orange-red base, feet olive green. JUVENILE: Crown bright rufous or rufous barred bronze, back and rump barred rufous and greenish bronze, wing brown with rufous edge of flight feathers with the coverts barred green or brown, tail T1 bronze-green with rufous notches on edge, other tail feathers dark with white tip and black bars,T5 barred white and rufous; chin to belly and under tail coverts white with bold but irregularly brown bars; iris dark brown, bill dark brown, feet gray. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, BPBM, FMNH, MVZ, MZB, SMF, TISTR, USNM, UWBM,WFVZ,YPM, ZRC.

Subspecies Chrysococcyx xanthorhynchus xanthorhynchus (Horsfield, 1821); male, head and upperparts glossy violet; NE India, Bangladesh, Yunnan, Burma, Thailand, Laos, Vietnam, Bay of Bengal islands (Andamans,

Violet Cuckoo Chrysococcyx xanthorhynchus 389 Nicobar Is), Malay Peninsula, the Greater Sundas (Sumatra, Java, Borneo), the Philippines (Palawan, Basilan); Chrysococcyx xanthorhynchus amethystinus (Vigors, 1831); male plumage glossy blue violet, less reddish violet; the Philippines (Cebu, Luzon, Mindoro, Samar, Mindanao, Catanduanes). Taxonomic notes: Cuculus malayanus Raffles, 1822, was based on a female of this species and is a synonym of C. xanthorhynchus (Oates 1887, Parker 1981), and older museum collections have specimens of female C. xanthorhynchus that were labeled “malayanus” because of the earlier confusion in names. A form C. x. bangueyensis Chasen and Kloss 1929 from Banguey (Banggi I) was described as having less white on the tail of both sexes, but the amount of white varies in populations in the Malay Peninsula and in Sumatra (Chasen and Kloss 1930, Medway and Wells 1976, MZB, ZRC). Plumage variation in the small glossy cuckoos of southeast Asia was confused in earlier works. Wells (1999) stated that juvenile C. xanthorhynchus has a black-streaked crown: this error traces back to Baker and Hume. Baker (1906, 1927) described the young of both C. maculatus and C. xanthorhynchus as streaked on the head, after Hume (1875) who described juvenile C. xanthorhynchus in Pegu, Burma, as streaked on the head. The bird described by Hume was rufous and appears to have been a juvenile Plaintive Cuckoo Cacomantis merulinus querulus, the only cuckoo in India and Burma with a streak-headed juvenile.

Measurements and weights C. x. xanthorhynchus: Wing, M (n ⫽ 20) 93–104 (100.2 ⫾ 2.8), F (n ⫽ 18) 93–106 (99.6 ⫾ 4.2); tail, M 63–73 (68.1 ⫾ 3.9), F 58–76 (67.8 ⫾ 4.8); bill, M 14.2–16.7 (15.6 ⫾ 0.7), F 13.7–17.5 (15.4 ⫾ 1.1); tarsus, M 12.4–15.6 (13.6 ⫾ 0.9), F 13.1–16.5 (14.4 ⫾ 0.9) (AMNH, ANSP, MCZ, SMF, WFVZ, YPM, ZMUC); C. x. amethystinus: Wing, M (n ⫽ 6) 97–101 (99.5 ⫾ 2.2), F (n ⫽ 5) 97–106 (101.4 ⫾ 3.6) (AMNH, CM, DMNH, FMNH, SMF, ZMUC). Weight, C. x. xanthorhynchus: M (n ⫽ 4) 18.7–26.4 (20.8), F (n ⫽ 4) 21.0–26.4 (22.8) (ANSP,

BMNH,WFVZ, ZRC); autumn migrants in Malay Peninsula, M (n ⫽ 10) 19.9–24.1, F (n ⫽ 7) 20.5–25.6 (Wells 1999), M (n ⫽ 1) 17.0 (ANSP); C. x. amethystinus: F (n ⫽ 1) 21.2 (FMNH). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⱖ 1 0 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 16 cm. Small cuckoo with upperparts glossy violet (blue in the Philippines except violet in Palawan and Basilan), barred underparts and a red bill; female has gray-brown head and nape (rufous in C. maculatus) and white speckling on face and a brown bill, red at the base; juveniles have the throat white barred black (same pattern but washed rufous in C. maculatus). Female differs from C. minutillus by the finer barring on chin and throat, a mostly rufous tail, a coppery crown and a red not yellowish eye-ring. In flight, male is dark, female has brown under wing with a weak white wing stripe.

Voice Calls, (1) a shrill descending trill, sometimes preceded by a triple note “seer-se-seer, seeseeseesee”, becoming louder and more rapid near the end; and (2) a loud “che-wick” or “kie-vik” at rate of once a second, given in flight, when two males are near each other, and in courtship (Whitehead, in Sharpe 1890, King and Dickinson 1975, Medway and Wells 1976, Lekagul and Round 1991, Smythies 1981, MacKinnon and Phillipps 1993,Wells 1999, Kennedy et al. 2000, Sheldon et al. 2001). Calls recorded (Scharringa 1999) match these descriptions: (1) four whistles around 4 kHz, “come, here my dear”, the first note held on one pitch, the second rising, the third rising higher and the fourth falling, and (2) pairs of short notes, a shrill “pe-a”, the first louder than the second, at 4.5 kHz and the pair of notes lasting less than 0.2 sec.

Range and status Southern Asia from the Himalayan foothills of NE India, Bangladesh, Bhutan and south China to southeast Asia (Burma, Thailand, Laos, Cambodia, Vietnam, Malay Peninsula), the Greater Sunda

390 Violet Cuckoo Chrysococcyx xanthorhynchus (Singh 1994), in Assam they tend to be at higher altitudes than Asian Emerald Cuckoo C. maculatus (Stevens 1915, Baker 1934), in Malay Peninsula they occur from coastal plains to slopes at 500 m (Wells 1999), in Java from sea level to 800 m (Hoogerwerf 1948, Grantham 2000), and in the Philippines usually below 1000 m (Kennedy et al. 2000). In nonbreeding season, several birds may feed near each other on outbreaks of caterpillars and aphids (Baker 1927). In feeding they perch motionless, creep up and down branches taking small insects, and sally flycatch (Smythies 1981).

Food Archipelago and the Philippines. Northern populations are migratory, appearing in summer from MayAugust; and southern populations are resident. Uncommon in the Andaman Is (Tidaker 1984, Grimmett et al. 1999). In breeding season, the cuckoos are widespread but scarce in China, northern India, Bangladesh and southeast Asia (Engelbach 1932, Smythies 1986, Cheng 1991, Round 1988, Thompson et al. 1993, Robson 2000a, Thomas and Poole 2003), Sumatra (Meyer de Schauensee and Ripley 1939, van Marle and Voous 1988) and Sarawak (Thewlis et al. 1996), locally common in Sabah (Smythies 1981), rare in Kalimantan (Holmes and Burton 1987, van Balen and Prentice 1997), and uncommon in the Philippines (Gilliard 1950, Kennedy et al. 2000). Migrants come to floodlights at night on Fraser’s Hill, in the Main Range in the Malay Peninsula from August through December, and migrants are recorded in the Malacca Straits during this period; a bird was attracted to a light in Sabah in April (Sheldon et al. 2001). Vulnerable to loss of forest (Round 1988), the cuckoos sometimes use disturbed patches of forest on slopes, logged sites and second growth (Wells 1999).

Habitat and general habits Forests, in the canopy of secondary evergreen forests, forest edge, thick bamboo forest, mangroves, Albizia and overgrown rubber and cocoa plantations, orchards and gardens. They occur mainly in the plains, in lowlands and lower hill slopes. In Arunachal Pradesh they live from 150 m to 2400 m

Insects, including caterpillars, flies, ants, beetles; and fruit (Smythies 1981, 1986, Gonzales 1983, Sody 1989).

Displays and breeding behavior Males occur in solitary territories, advertise with long circling song flights over the forest canopy, and less often they sing from high perches in the canopy of tall trees. Males countersing (Wells 1999).

Breeding A fledged young cuckoo was fed by sunbirds in late May or early June in the Malay Peninsula (Wells 1999), an egg perhaps of this species was taken in February in Java (Hellebrekers and Hoogerwerf 1967), a laying female was taken in April in Sabah (Sheldon et al. 2001, WFVZ 37582), and a laying female had an oviduct egg in April on Mt Isarog in Luzon (Goodman and Gonzales 1990, FMNH 265505). Brood-parasitic. In India the hosts are thought to be Crimson Sunbird Aethopyga siparaja and Little Spiderhunter Arachnothera longirostra, the same as for C. maculatus (Baker 1927, Harrison 1969, Becking 1981). On Singapore a fledged young cuckoo was fed by a pair of Plain-throated (Brown-throated) Sunbirds Anthreptes malacensis (Wells 1999). In Java, an egg, perhaps this cuckoo’s, was in a nest of Yellow-bellied Prinia Prinia flaviventris (Hellebrekers and Hoogerwerf 1967). Baker (1906) proposed other hosts particularly small babblers as hosts of this cuckoo, based on cuckoo eggs in babbler nests, but

Diederik Cuckoo Chrysococcyx caprius 391 later made no reference to these records when he attributed the eggs to Banded Bay Cuckoo Cacomantis sonneratii (Baker 1927, 1934). Eggs (identified in India by a logic of elimination in their local distribution from other glossy cuckoo species) are less glossy and more rounded than C. maculatus, whitish buff or pink speckled and blotched red or violet and with more olive brown secondary markings, 16.4 ⫻ 12.3 mm (Baker 1906, 1927, 1934, 1942,

Inglis 1908).An oviduct egg was off-white with violet tint and small (1–2 mm) black and white splotches distributed evenly over the surface; the egg was broken and not measured (Sheldon et al. 2001). No observations of identified specimens of feathered juveniles in the care of their foster parents are known, and the eggs seen in nests are uncertain in identification, much as for C. maculatus.The incubation and nestling periods are unknown.

Diederik Cuckoo Chrysococcyx caprius (Boddaert, 1783) Cuculus caprius Boddaert, 1783, Table des Planches enluminéez d’histoire naturelle, p. 40. (Cape of Good Hope) Other common names: Didric Cuckoo,Whitebellied Didric Cuckoo. Monotypic.

Description ADULT: Male, head to back glossy bronze-green, wing coverts glossy bronze-green with large white spots, wing blackish with white bars across inner vane of primaries, tail dark green with white tips and white notches and spots along both edges of rectrices (T1 lacks white spots), upper wing coverts with white spots, face with white forehead streak and white eye streak; underparts white with a thin green malar streak, flanks, thighs and under tail coverts barred green, under wing coverts and flight feathers barred black and white, tail blackish with broad white spots from below; eye-ring red, iris red, bill black, feet gray. Female, like male but duller, with or without the forehead streak; some females are green above, others rufous or intermediate in plumage color, the upper wing coverts with white to buff or rufous spots, tail dark green with rufous notches and spots and T2 to T4 with broad rufous bars; underparts white to buff, throat and breast usually washed brown in southern Africa, breast often lightly streaked, tail black and rufous with broad white bars below; eye-ring brown, iris hazel brown to tan or gray, some tan or gray with dark flecks, some brown with a yellow outer ring, and others yellow centrally with brown flecks and a

brown outer ring, not red (in southern and East Africa), bill black. JUVENILE: Sexes alike, above plumage dull green or bright rufous (or both, and intermediate plumages with barred green and rufous are common), some birds rufous only on crown, upper wing coverts with pale spots (except in rufousphase juveniles); underparts white, throat with dark streaks, breast and belly with blackish or dark green spots (not bars or streaks); iris gray to brown, bill red. NESTLING: Naked at hatching, skin pink darkening to blackish; the eye-ring becomes red while birds are in juvenile plumage; iris gray changing with age to brown; bill orange to red, gape red, edge of gape yellow. SOURCES: AMNH, BMNH, BWYO, CM, CU, DMNH, FMNH, LSU, MCZ, MVZ, ROM, UMMZ, USNM, YPM, ZMS.

Geographic variation West African cuckoos are smaller on average and have been described as distinct, and in recent years Clancey (1990) has recognized a subspecies C. c. chrysochlorus (Cabanis and Heine, 1863). Geographic variation in size is continuous from West to East and southern Africa, and the overlap in size between West and East Africa is considerable, as noted also by Friedmann and Loveridge (1937), and no subspecies are recognized here.

392 Diederik Cuckoo Chrysococcyx caprius Females of different egg morphs are similar in appearance. In the Eastern Cape, females with oviduct hard blue eggs with brown spots did not differ in body size (UMMZ 216741, 216743, 216763; wing 119, 120, 123), plumage color or iris color from females with hard white eggs with small brown flecks (UMMZ 216740, 216746; wing 121, 124).A female with a blue egg and brown spots was near a weaver colony, and a female with a white egg with blackish flecks was chased by Cape Sparrow Passer melanurus near the sparrow nest. In northern Nigeria a female (UMMZ 216736; wing 108) with a hard unmarked white egg was taken near a colony of Village Weaver Ploceus cucullatus and did not differ in appearance from other West African females of this species. Rufous-phase juveniles occur in areas where cuckoos are reared by seed-eating bishops Euplectes species; normal green-phase juveniles occur in areas where the common hosts are Ploceus weavers.

Measurements and weights West Africa (Senegal to Cameroon): Wing, M (n ⫽ 13) 102–117 (111.4 ⫾ 4.7), F (n ⫽ 30) 103–116 (112.0 ⫾ 3.0) (AMNH, CMNH, LSU, MCZ, UMMZ, YPM); Uganda: Wing, M (n ⫽ 10) 108–119 (112.2 ⫾ 3.4), F (n ⫽ 10) 106–118 (114.0 ⫾ 5.0) (FMNH); Malawi:Wing, M (n ⫽ 98) 109–124 (116.6 ⫾ 3.2), F (n ⫽ 54) 110–125 (117.5 ⫾ 3.4) (Hanmer 1995); Eastern Cape: Wing, M (n ⫽ 6) 114–120 (116.5 ⫾ 2.2), F (n ⫽ 23) 117–128 (121.3 ⫾ 2.9); tail, M 82–92 (85.3 ⫾ 4.2), F 87–98 (90.7 ⫾ 2.8); bill, M 15.3–17.6 (16.4 ⫾ 1.0), F 16.6–19.9 (17.7 ⫾ 0.9); tarsus, M 17.3–20.4 (18.3 ⫾ 1.2), F 17.4–20.0 (18.3 ⫾ 0.8) (UMMZ). Weight, Nigeria: M (n ⫽ 5) 27.0–29.0 (24.5), F (n ⫽ 7) 24.5–31.0 (30.2) (DMNH); Malawi: M (n ⫽ 102) 20.5–37.1 (29.3), F (n ⫽ 64) 28.3–44.3 (36.7) (Hanmer 1995); Eastern Cape: M (n ⫽ 5) 30.8–38.0 (34.3), F (n ⫽ 9) 36.0–46.1 (40.9) (UMMZ). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 1 0 ⬎ 2 ⬎ 1.

Field characters Overall length 20 cm. Males are glossy green above with a white eye streak, white spots on wing, white

below with green bars on flanks, and a red iris; females green to rufous above with whitish spots on wing, below white with dark streaks or bars on breast, bars on belly and flanks, and brown or gray iris. Wings are entirely barred below, in flight (in Klaas’s Cuckoo C. klaas, male wing lining is nearly white, female wing lining is white with narrow dark bars). Juveniles have streaked and spotted throat and breast and red bill.

Molt Molt is gradual and lasts for several months. In juveniles the primary molt begins at about four months of age (Hanmer 1995). Molt is often incomplete in the primaries, secondaries and coverts in first-year birds.

Voice Male has clear emphatic whistle, the first notes rising in pitch, “dee-dee-dee-diederik”! Rapid and high-pitched “diederik” calls are also given in aerial chases between males. Male gives a wavering whinny-whisper “weahweahweahweah” (“caterpillar call”) in display with a caterpillar to a female. The female calls “deah, deah, deah, DEAH” to a male and he brings her a caterpillar (Chappuis 1974, 2000, Stjernstedt 1993, RBP). Songs are remarkably the same across the range of the species.

Range and status Africa south of the Sahara, and in southern Arabian Peninsula (Yemen, Oman) (Irwin 1988, Short et al. 1990, Porter 1996). Resident in low latitude tropics, and a migrant within Africa, where seasonal in the north and in the south, and in the drier areas during the rainy season. In Senegal it is a seasonal visitor during the rains in the north but present all year in the south (Morel and Morel 1990); in The Gambia from June to October, rarely later (Gore 1990, Barlow et al. 1997); in Ghana mainly in the wet season, and a seasonal visitor in the north (Greig-Smith 1976, Grimes 1987); in the inundation zone of the Niger River from June to November (Bannerman 1951); in Togo and southern Benin they are locally resident in the south and seasonal visitors in the northern woodlands from April to August (Cheke and Walsh 1996, Anciaux 2000); in northern Nigeria they occur during the

Diederik Cuckoo Chrysococcyx caprius 393 regions near colony-nesting ploceids. Annual adult survival is c. 59% (Hanmer 1995).

Habitat and general habits

rains from May to October then migrate south for the dry season (Elgood et al. 1994). In Cameroon they occur throughout the rains but are absent north of 6°N in the dry season November to March (Louette 1981a, Quantrill and Quantrill 1998). In Gabon they appear in savannas as nonbreeding visitors from the Sahel in January and February, and in all seasons they are residents around weaver colonies (Christy and Clarke 1994); in forest clearings they occur all year, most are seen from March to September (Brosset and Erard 1986); resident on Gulf of Guinea islands (Pérez de Val et al. 1997).They are seen en route in northern Somalia in March to early June and again in September (Ash and Miskell 1998); in Sudan the cuckoos are breeding visitors during the rains but resident in the extreme south (Lynes 1925, Nikolaus 1987). In East Africa they are resident in wet areas and wet-season visitors in dry areas ( Jackson 1938, Lewis and Pomeroy 1989, Zimmerman et al. 1996), in Zanzibar and Pemba islands they breed from October to February, with the November-December monsoons and not the March-May monsoons (Pakenham 1979). In Malawi, most are present and sing from October to March; a few appear in other months as well (Benson and Benson 1977); in Zambia the birds are seasonal from October to April, and they are seasonal during the rains in southern Africa (Vernon et al. 1997). Seasonal in Dhofar, Oman and the southern Arabian Peninsula where they visit in summer from May to October (Gallagher and Woodcock 1980, Brooks et al. 1987, Porter et al. 1996).Accidental in Cyprus (Lobb 1983) and at Eilat, Israel (van den Berg 1994). Common in many

Semi-arid thorn scrub, acacia savanna, open woodlands, and edge of marsh habitat. In the forest zone they occur in open woodland mosaics on inselbergs and trees of open swamps along forest rivers, and in large clearings and rice fields; in semi-arid regions they occur near water, often by weaver colonies, and near human cultivation where they parasitize weavers (Skead 1952, Gatter 1977, Vernon et al. 1997). Generally they occur below 1200 m, and not on higher mountain masses above the forest belt.The feeding bird moves actively from perch to perch several times a minute, or waits a few minutes then flies to a caterpillar and grabs it, or peers from a perch then sallies out and grabs an insect from the ground or trunk of a tree, and returns to its perch with the insect. It takes caterpillars from the ground or low herbs (lucerne, alfalfa) by hovering over the plants, and grabs insect in its bill as it is airborne. Hairy caterpillars are killed and wiped on a perch or shaken in its bill, the caterpillar grasped near the head, the guts slung out of the end while hairs remain on the insect when the cuckoo swallows it. After it leaves the nest, a fledgling fed by a foster parent may feed itself as well, taking a caterpillar and skinning it by holding the insect at one end, flicking it until the skin separates from the body, then shaking its head so that the caterpillar skin flies off, and eating the insect (Gargett and Webb 1973).

Food Insects, mainly caterpillars, also grasshoppers (Irwin 1988), termites (workers of Hodotermes mossambicus, RBP), alates taken on the wing, the termites’ wings removed when the bird rubs them off against a perch (Skead 1995), beetles; females take eggs from host nests ( Jensen and Jensen 1969, Payne 1973a, Macdonald 1980).

Displays and breeding behavior The male calls in bushes or trees where he attracts females, courting them on repeated trips with a caterpillar which he feeds to them. He also courtship feeds

394 Diederik Cuckoo Chrysococcyx caprius females at other times of day, as in the hour after sunrise. Males display near a tree with a weaver colony, or away from active nests. Each male has a site where he regularly displays, day after day, in the early morning and again for an hour in late afternoon. Several females visit a male at this site. Courtship feeding (Color Plate 11, 6a–c) is traced from a video taken by Ole Post in The Gambia. In courtship feeding male flies to female with his wings stiff, fluttering and not rising above horizontal, calling as he flies, and when he perches with a caterpillar, a whinny-whisper “weahweahweahweah”. Female calls to male a loud “deah, deah, deah, DEAH”, rising in pitch and amplitude, sometimes gives a series of notes each phrase with a rising whistle then a higher whistle that drops in pitch (RBP, 13 Nov 1965). Other male-female feedings are silent. One male fed a female 15 times in 30 min in Zimbabwe (Erwee 1989). In a pair that I watched near Magadi, Kenya, for an hour after sunrise (31 May 1967), the male fed a female 12 times in 30 min while they remained within an area of 100 ⫻ 400 m. Male perches near and faces her, leans forward arcing towards her, his tail spread and held down, bobs his head and neck about 8 times in three sec, moving up and down, or swinging head from side to side, and appears to force the food upon the female; she leans back then forwards as she takes the caterpillar, wipes it on her perch, then swallows it. In copulation either after courtship feeding, or when feeding is not observed, the female raises and fans her tail, and the male mounts her from behind. Some females that accept a caterpillar are ovulating (ovum entering the oviduct, ready to be fertilized), some have a soft-shelled egg, or a hard-shelled egg ready to lay, and others are not breeding (no large eggs in ovary) (UMMZ). Females call and are fed by a male at any stage of their laying cycle (RBP). Females that parasitize different species of hosts may mate with the same male (Verheyen 1953). Territorial, male chases away other males from an area and sings prominently in the area. In aggressive display, male chases another male, calls a rapid version of the “diederik” song, higher in pitch than in the advertising song. Male-male chases are common in the breeding season; they fly with rapid deep wingbeats, sometimes for 30 min or longer. Male pursues female in a rapid twisting flight as far as 400 m. On

the other hand in some areas males are not obviously territorial and several singing males use the same area around a colony of hosts (Macdonald 1980). Fledged young with a red bill are sometimes fed by the adult male (this is much less commonly observed than male feeding female) in misdirected courtship feeding (Ottow and Duve 1965, Friedmann 1968, Skead 1995). Male gives a flutter flight towards the young as he arrives with a caterpillar, offers it with the same flight and perch displays and calls that he gives to a female, although the bill color and plumage differ from that of the adult females. Males feed both male and female fledgling cuckoos (UMMZ). Female lays a set of 3 eggs on alternate days, then pauses a few days between sets. A female lays c. 20 eggs during a breeding season. Females breed in their first year (when they have a few retained juvenile feathers) as well as in their later adult years (Payne 1973a, 1974). A female observes from a perch near the host nests, then flies to the nest alone. She visits the host nest alone, peers inside, enters when the host is absent, is chased away when the host is present and is pursued by the nesting weaver for at least 100 m. Female takes host egg from nest, flies up to 100 m to a perch where she eats the egg.The visits sometimes involve taking an egg and returning to lay her own egg. A female visit to a weaver nest takes less than 10 sec when she removes an egg. One female went into a nest of Lesser Masked Weaver Ploceus intermedius, then laid in the nest, when she became stuck in the nest entrance and wasps stung the struggling cuckoo to death ( Jensen 1975). More than one female may lay in the same weaver colony, as indicated by variation in the color of cuckoo eggs in the nests in a colony (Hunter 1961, Friedmann 1968, Collias and Collias 1970) and by observation of individually distinctive female cuckoos in a weaver colony (Macdonald 1980). Nests in large colonies of Red Bishop Euplectes orix are less likely to be parasitized than nests in small colonies (Ferguson 1994, Lawes and Kirkman 1996).

Breeding With the rains, in northern Senegal they lay from August to October, in The Gambia breed from July

Klaas’s Cuckoo Chrysococcyx klaas 395 to September, in western Mali in September and October (Bruggers and Bortoli 1979, Lamarche 1980, Barlow et al. 1997), in rice fields in lowland Liberia they sing all year with peaks in May-June and September-October (Gatter 1997), in coastal Ghana they breed from March to July (Macdonald 1980, Grimes 1987), in northern Nigeria mainly from April to September (Elgood et al. 1994), in Bioko in October (Eisentraut 1973), in Malawi from October to April (Benson and Benson 1977, Hanmer 1995), in Kenya at Lake Baringo in both rainy seasons (short rains April-May and long rains NovemberJanuary) (Brown and Britton 1980, Zimmerman et al. 1996), in Angola from December-June (Dean 2000), in Botswana and Zimbabwe from October to March (Irwin 1981, Skinner 1996), in South Africa from late October to mid January in the Eastern Cape, a month or two later in the north (Rowan 1983, Maclean 1993,Vernon et al. 1997). Brood-parasitic. Hosts, mainly weavers Ploceus (Spectacled Weaver P. ocularis, Black-necked Weaver P. nigerrimus, Masked Weaver P. velatus, Village Weaver P. cucullatus, Cape Weaver P. capensis, several other weavers both colonial and solitary) bishops Euplectes (especially Red Bishop E. orix), and Cape Sparrow Passer melanurus; others are scrub-robins Cercotrichas (Karoo Scrub-robin C. coryphaeus, Kalahari Scrub-robin C. paenia, Red-backed Scrubrobin C. leucophrys, Rufous-tailed Scrub-robin C. galactotes), prinias (Karoo Prinia Prinia maculosa), warblers (Little Rush Warbler Bradypterus baboecala), sunbirds (Red-chested Sunbird Nectarinia erythro-

cerca,Tring) and wagtails Motacilla spp. (Chapin 1939, Friedmann 1948, 1956, 1968, Skead 1947, 1952, 1995, Reed 1953, Rowan and Broekhuysen 1962, Ottow and Duve 1965, Payne and Payne 1967, Jensen and Vernon 1970, Kemp et al. 1972, Macdonald 1980, Rowan 1983, Brosset and Erard 1986, Pérez de Val 1996, Jackson 1998, Squelch and SafeSquelch 1994, Lawes and Kirkman 1996, Leonard 1998b). Within a single locality the cuckoos parasitize more than one species of hosts, including several species of weavers, bishops Euplectes, Cape Sparrow and others (West et al. 1964, Reed 1968). Eggs are variable in color, white, greenish or blue, and spotted or unspotted, often matching the color and pattern of the host species (Reed 1953, 1968, Hunter 1961, Ottow and Duve 1965, Markus 1967, Payne 1967, Friedmann 1968, James 1970, Jensen and Vernon 1970, Jensen 1975, ColebrookRobjent 1977, 1984, Lawes and Kirkman 1996, Tarboton 2001); 21 ⫻ 15 mm. The incubation period is 12 days. The nestling cuckoo is 2.3 g on the day of hatching. It evicts the host eggs and young on day 2; the evicting response occurs until day 4; the eyes open on c. day 10. The young cuckoo fledges in 20–22 days at 35 g and remains with its foster parents for 3 more weeks (Reed 1968, Chalton 1976, 1991, Skead 1995). Young in nests of bishops are reared on seeds late in the nestling period.The young cuckoos in nests of certain host species (Masked Weaver, Red Bishops and Cape Sparrow) differ in their begging calls (Reed 1968).

Klaas’s Cuckoo Chrysococcyx klaas (Stephens, 1815) Cuculus klaas Stephens, 1815, in Shaw’s General Zoology, 9, p. 128. (Cape Colony) Monotypic.

Description ADULT: Male, upperparts and side of neck glossy green, wing coverts green, primaries barred black and white, the outer vane glossy green, inner secondaries glossy green, upper tail coverts green with white outer vane, tail green, the inner rectrices

unmarked, outer rectrices white with large green subterminal spot on outer vane and incomplete black bars mainly near tip, green face extends down side of neck and breast, small white patch behind eye; underparts white, green half collar extends in front of wing to side of breast, flanks narrowly barred green, legs white with feathered green streaks, under wing coverts white with fine black bars, under wing flight feathers with broad bars and black secondaries contrast with white

396 Klaas’s Cuckoo Chrysococcyx klaas wing coverts, tail nearly white from below; eyering green, iris dark brown, bill green, feet greenish. Female, head and upper back gray-brown, lower back barred green and brown, wing coverts barred green and brown, wing brown with light rufous notches on outer vane and light rufous bars on inner vane, rump coppery brown, outer vane of upper tail coverts white, base of tail white, T1 and T2 coppery brown,T3 to T5 white with large coppery subterminal spot on outer vane and 1–2 incomplete black bars, face grayish brown extending down side of neck and breast, small white patch behind eye; underparts white finely barred brown on breast and flanks, the bars narrow and broadly V-shaped in southern Africa, legs with coppery feathered pants, under tail coverts white with coppery subterminal V’s, under wing coverts white with narrow dark bars; eye-ring paler green than in male, iris light brownish gray, bill greenish with black tip. JUVENILE: Sexes alike, head and back barred green and buff, rump barred dark brown and buff, nape feathers with white base, face with indistinct white streak behind the eye, wing barred green and buff, the green replaced by dark brown on inner vane of flight feathers, central tail feathers bronze-brown with partial buff bars near base and edge, outer tail feathers white with three to five partial or complete black bars near the tip, upper tail coverts green medially with a white outer vane, rump white viewed from the side; underparts white narrowly barred dull bronze-green from chin to under tail coverts, under wing coverts white barred blackish; iris hazel, bill black. NESTLING: Naked at hatching, skin brown, darkens to deep blackish olive; iris gray-brown, bill dark horn, gape orange. SOURCES: AMNH, BMNH, BWYO, CM, FMNH, LSU, MCZ, MVZ, ROM, SMF, UMMZ, YPM, ZMUC.

Molt In Malawi mainly March-July, some birds complete the molt by May. Feathers are often retained from

juvenile to first adult plumage, especially the wing coverts and remiges. Juvenile primary molt begins at 4 months and is completed in 75–80 days.

Measurements and weights Cameroon: Wing, M (n ⫽ 11) 90–102 (96.0 ⫾ 4.3), F (n ⫽ 6) 92–97 (94.7 ⫾ 1.9); tail, M 67–74 (69.5 ⫾ 2.3), F 63–71 (67.8 ⫾ 3.1); bill, M 13.1–1 6.8 (14.7 ⫾ 0.9), F 14.4–16.6 (15.0 ⫾ 0.9); tarsus, M 13.7–16.1 (14.7 ⫾ 0.7), F 14.6–15.4 (14.9 ⫾ 0.4) (MCZ, FMNH); Kenya: Wing, M (n ⫽ 13) 92–101 (96.9 ⫾ 2.3), F (n ⫽ 10) 93–103 (97.4 ⫾ 3.9); tail, M 61–72 (69.5 ⫾ 7.7), F 63–72 (67.4 ⫾ 2.8); bill, M 12.1– 15.4 (13.4 ⫾ 1.0), F 12.9–15.4 (14.0 ⫾ 0.8); tarsus, M 13.0–15.1 (14.2 ⫾ 0.8), F 13.1–15.4 (14.2 ⫾ 0.7) (FMNH); South Africa: Wing, M (n ⫽ 10) 96–104 (99.5 ⫾ 2.4), F (n ⫽ 9) 95–105 (98.9 ⫾ 5.9); tail, M 67–80 (73.4 ⫾ 4.1), F 73–83 (76.0 ⫾ 3.2) (BMNH, UMMZ). Weight, M (n ⫽ 30) 22.7–32.8 (26.2], F (n ⫽ 38) 24.7–35.1 (28.6) (Hanmer 1995). Wing formula, P9 ⱖ 8 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10. Geographic variation:These cuckoos are larger on average in south-central and southern Africa than in East and West Africa.The largest birds in Cameroon (wing 102) were taken in October and January, when the species would breed in southern Africa, the birds resident rather than non-breeding migrants. Females in southern Africa are unlike males in plumage, whereas females in East Africa have a darker gray head and breast (“a shadow version of a male”), and females are variable in West Africa (Guinea to Cameroon and Central African Republic), some like East African females while others have nearly male plumage: glossy green above with no bars on the wing and with the coverts all green, throat white, belly unbarred, usually barred on flanks (males with no visible bars), or glossy copper with bars on flanks and wing and with an open gray vest. The sex of two females in male-like plumage was confirmed through an egg in the oviduct; a third was breeding and had an ovum of 6 mm (Serle 1965, Irwin 1987b, LSU 148160 in Liberia, AMNH 297396 in Bioko, FMNH 832029 in Central African Republic).

Klaas’s Cuckoo Chrysococcyx klaas 397

Field characters Overall length 18 cm. Small glossy green cuckoo, male with a white patch behind the eye, green half collar extends onto sides of breast, female with a whitish streak behind eye and indistinct half collar on side, gray throat and underparts finely barred. Both sexes have unspotted wing, white outer tail feathers and outer upper tail coverts, and the underwing has white coverts and barred primaries (underwing is all barred in Diederik Cuckoo C. caprius). Juvenile Klaas’s Cuckoo has crown barred green and brown to buff ( juvenile African Emerald Cuckoo C. cupreus has white bars on crown), above green with buff not rufous bars, rump buff and dark brown with a white edge, underparts with thin bronze bars (not broad green bars), and more black bars on white outer tail feathers.

Voice Song, two clear whistles,“who is he?” or “meit-jie”, the first phrase slurred up, the second slurred down, often given three songs in a row, male sings on perch for as long as 20 min (Chappuis 1974, 2000, Stjernstedt 1993, Zimmerman et al. 1996, Barlow et al. 1997). In chase or in response to playback of song, male gives a harsh chatter on high pitch, “cheecheechee” (RBP, Amanzi, Eastern Cape, 20 Nov 1966).Also a “quip quip” then a trill by female while male gives a whistle in courtship when feeding a caterpillar to the female (Stjernstedt 1998). Fledged young begging for food from its foster parent (Cape Crombek Sylvietta rufescens) gives a high thin ascending trill (unlike C. caprius), and a harsh call (RBP).

Range and status Africa from Senegal east to Sudan, Ethiopia and South Africa; southern Arabian Peninsula, Bioko. Resident in equatorial areas, seasonal migrant at higher latitudes, arriving during the rains, leaving afterwards, though some birds remain in nonbreeding season. In northern Senegal they occur only in the rainy season (Morel and Morel 1990), in Ghana they are mainly wet-season visitors (Grimes 1987), in Togo they occur in north during rains and in south throughout the year (Cheke and Walsh 1996),

X

at Kumba, Mt Kupé and Yaounde, Cameroon, they occur and sing in all months (Serle 1950b, Quantrill and Quantrill 1998, Bowden 2001), in parts of Sudan they appear seasonally with the rains (Nikolaus 1987). Gulf of Guinea, resident (Amadon 1953, Eisentraut 1973, Pérez de Val et al. 1997). In southern Zaire they are present from July to April (Verheyen 1953), in Zambia there are records in all months (Aspinwall and Beel 1998), and in they are Zimbabwe migrant although a few are around and sing in July (Sheehan 1999). In Yemen they visit in late spring and summer (Brooks et al. 1987, Porter et al. 1996). Occasional in Arabia, where cuckoos may be rare breeding residents to the Tihama ( Jennings 1981). Resident and sing in winter in Cape Town region. Common, they persist in residential areas when flowers are available for their sunbird hosts.They feed with other species in mixed foraging flocks.

Habitat and general habits Open woodland with Combretum and Terminalia, miombo with Brachystegia and Julbernardia, mopane woodland with Colophospermum mopane, dense acacia, forest, forest edges, thickets, savanna, gardens (Chapin 1939, Greig-Smith 1976, Irwin 1988); in West Africa in lowlands to 800 m (Gatter 1997), in East Africa lowlands to 3000 m, most records (84%) are within the 500⫹ mm rainfall region (Lewis and Pomeroy 1989). They forage in foliage, less often they flycatch. They feed actively, peer about, twist the head and neck to peer into foliage and capture an insect within reach, or they fly or hop to the

398 Klaas’s Cuckoo Chrysococcyx klaas prey. The cuckoo holds the head of a caterpillar in its bill, shakes the caterpillar from side to side, and hits larger caterpillars against a branch. Male often alternately feeds and sings, 10–60 sec between captures (RBP).

Food Insects, mainly caterpillars, and butterflies, bug nymphs, cotton-stainer bugs Dysdercus intermedius, stink bugs, beetles (taken from surface of bark), small orthoptera, termites (Benson 1942, Irwin 1988, Skead 1995, UMMZ).

Displays and breeding behavior In courtship behavior, male offers caterpillar to a female (Friedmann 1968, Leonard 1998b, Stjernstedt 1998, Salewski 2000). Male bobs up and down in front of female holding the caterpillar in his bill, sways from side to side with spread wings and tail, lunges forward and to one side, then back, then forwards and to the other side. Male and female switch their tails vigorously up and to the side. After the male presents a caterpillar to her, the female wipes the caterpillar on the branch (RBP, Bathurst Forest, Eastern Cape, 26 Oct 1967). Male gives a rising whistle “meit”, while female has a long whistle “whoooeeeoo” that rises then falls in pitch, lower in pitch than the male song (RBP, 21 Oct 1966, Alexandria Forest, Eastern Cape; Merensky Reserve, Transvaal, 21 Dec 1966; Magadi, Kenya, 12 June 1967) and sometimes gives a short chirp. After he feeds her, female remains on the feeding site, male flies out, finds another caterpillar, returns to her and feeds her again.The male feeds her as many as 10 caterpillars in 10 minutes (Leonard 1998b). Females that take a caterpillar often have an ovum ready to be fertilized, and one (RBP 3814) at Bedford, Eastern Cape, had a large yolky ovum attached to the ovary and engulfed by the mouth of the oviduct. Fledged short-tailed young cuckoos are occasionally fed by an adult male (Moreau and Moreau 1939), the fledgling perhaps mistaken for a breeding female.

Breeding and life cycle During the rains: in Senegal and The Gambia they breed from September to November (Morel and Morel 1990), in Ivory Coast in February (courtship

feeding: Salewski 2000), in Nigeria March to September (Elgood et al. 1994), in Gabon they sing from September to February (Christy and Clarke 1994), in East Africa east of the rift valley breeding records are from December to June (Brown and Britton 1980), in Angola from September to December (Pinto 1983, Heinrich 1958, Dean 2000), in Malawi from October to February (Benson and Benson 1977), in southern Zaire from August to January (Lynes 1933, Verheyen 1953), in Zambia from December to February (Benson et al. 1971), in Zimbabwe from September to March (peak October and November) (Irwin 1981), in South Africa in Natal from October to January (Dean 1971), in the Eastern Cape from October to December and occasionally as late as May (Skead 1971), and in the Cape region during winter rains from July to September (Rowan 1983).A female lays in sets of 3–4 eggs, with a day between eggs, on average about 24 eggs in a season (Payne 1973a, 1974). Brood-parasitic. Hosts, small insectivorous songbirds, mainly warblers (e.g., Bar-throated Apalis Apalis thoracica, Long-billed Crombek Sylvietta rufescens, Grey-backed Camaroptera Camaroptera brevicaudata, Yellow-bellied Eremomela Eremomela icteropygialis), flycatchers (e.g., Chin-spot Batis Batis molitor, Cape Batis B. capensis, Black-throated Wattleeye Platysteira peltata) and sunbirds (e.g., Copper Sunbird Nectarinia cuprea, Red-chested Sunbird N. erythrocerca, Scarlet-chested Sunbird N. senegalensis, Greater Double-collared Sunbird N. afra, Marico Sunbird N. mariquensis, Dusky Sunbird N. fusca), and African Pied Wagtail Motacilla aguimp (Braun 1934, Jackson 1938, Chapin 1939, Friedmann 1948, 1956, 1968, Skead 1952, 1954, 1995, Macleod and Hallack 1956, Payne and Payne 1967, Jensen and Jensen 1969, Jensen and Clinning 1974, Irwin 1981, Rowan 1983, Maclean 1993, Bishop et al. 1995, Steyn 1996, Payne 1998). The cuckoos are more selective in use of a host species within a local population than across a region. Eggs vary in color (white, greenish white, blue), spotted with brown or rufous, 19 ⫻ 13 mm ( Jensen and Clinning 1974, Tarboton 2001). An egg in the oviduct of female taken on 10 Oct 1965 at Committees, Eastern Cape, was greenish with small brown spots (UMMZ 216730). Incubation period is 11–12 days, nestling

Yellow-throated Cuckoo Chrysococcyx flavigularis 399 evicts eggs and young of host, nestling period is 19–21 days, and a fledged bird remained with its foster parents for 25 days after leaving the nest.

Begging call, a husky “tsseek,” repeated and increasing to a rapid stuttering with a wing quiver when the foster parent approaches (Skead 1995).

Yellow-throated Cuckoo Chrysococcyx flavigularis (Shelley, 1880) Chrysococcyx flavigularis Shelley, 1880, Proceedings of the Zoological Society of London, 1879, p. 679, pl. 50. (Elmina, Fantee [Ghana]). Other common names: Yellow-throated Green Cuckoo,Yellow-throated Glossy Cuckoo. Monotypic.

Description ADULT: Male, upperparts and sides of neck coppery brown washed bronze-green, wing coverts coppery green with inconspicuous and variable rufous bars, wing brown, tail T1 purplish bronze with light bronze tip, T2 bronze-brown with buff outer vane at base, a wide blackish subterminal bar and white tip,T3 with white outer vane and rufous inner vane, the inner vane with incomplete blackish bars, both vanes with a broad subterminal black bar and white tip, T4 white with bronze-brown basal third of inner web and a black subterminal bar,T5 white often with an incomplete narrow black subterminal bar near the tip (in sum, the three outer rectrices white with bronze-brown base and subterminal black bar); underparts, center of chin and throat bright yellow bordered by dark green, breast, belly and under tail coverts finely barred buff and greenish brown, wing barred buff on underside of inner edge of flight feathers, under wing coverts barred buff and greenish brown; eye-ring yellowish green, iris yellow, bill greenish yellow, feet yellow. Female upperparts similar to male, the face finely barred; underparts green barred buff from throat to under tail coverts; eye-ring greenish yellow, iris dark brown to pinkish brown, bill black, feet dull yellow. JUVENILE: Above, coppery brown, finely barred with buff from crown to rump, underparts finely barred as in adult female. Males molt directly from juvenile plumage to adult plumage.

NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, CM, FMNH, LACM, ZMB.

Measurements and weights Wing, M (n ⫽ 10) 91–98 (94.6 ⫾ 2.8), F (n ⫽ 8) 91–99 (94.8 ⫾ 2.4); tail, M 56–72 (65.0 ⫾ 6.0), F 56–69 (61.6 ⫾ 4.4); bill, M 14.7–18.2 (16.1 ⫾ 1.1), F 14.4–18.2 (17.3 ⫾ 1.6); tarsus, M 13.1–14.9 (13.8 ⫾ 0.7), F 12.3–15.0 (14.1 ⫾ 1.2) (AMNH, BMNH, CM, FMNH, LACM). Weight, M (n ⫽ 7) 27.5–31 (29.0), non-breeding F (n ⫽ 1) 30, laying F (n ⫽ 1) 30.5 (AMNH, LACM). Wing formula, P8 ⫽ 7 ⬎ 6 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 1 0 ⫽ 2 ⬎ 1. A subspecies Chrysococcyx flavigularis parkesi Dickerman 1994, in Central Africa (Cameroon, Zaire and Uganda), holotype a female LACM 70110 from Bwamba Forest, western Uganda, was described as distinct based on color of the underparts (belly and under tail coverts browner, less creamy white, than West African birds). Specimens show the same color of the underparts in females from Central and West Africa, including specimens at BMNH; and the holotype of C. flavigularis, a male from Ghana illustrated in Shelley’s species description, has the same color and pattern of the underparts as males from Uganda. As noted by Friedmann (1969) the under tail coverts are variable within Uganda. The subspecies parkesi cannot be recognized.

Field characters Overall length, 19 cm. Male dark bronzy copper above, broad yellow stripe on chin and throat, breast and belly buff finely barred greenish brown; female with face and underparts buff finely barred greenish brown. From below, tail is all white except for black subterminal bar, wing is dark, not barred as in female African Emerald Cuckoo C. cupreus.

400 African Emerald Cuckoo Chrysococcyx cupreus

Voice Song, a clear flutelike whistle of 9–12 notes on same pitch, first the longest, a plaintive note, then a series of notes accelerating then fading away, “tui, tiuuhuhuhuhuhuh” lasting about 3 sec. Both sexes give a two-note whistle (the second note lower), “chuwee-chwee”.A female approached, in response to a human whistle imitation of this call (Friedmann 1969, Brosset and Erard 1986, Chappuis 2000, Stevenson and Fanshawe 2002).

Range and status West and Central Africa in forest zone from Sierra Leone, Liberia, Ivory Coast and Ghana through Nigeria, southern Cameroon and Gabon to Zaire, S Sudan (one record) and southwestern Uganda (Bannerman 1933, Friedmann 1968, 1969, Nikolaus 1987, Irwin 1988, Christy and Clarke 1994, Elgood et al. 1994, Cheke and Walsh 1996, Gatter 1997, Stevenson and Fanshawe 2002). Resident, they sing all year. Rare in most of West Africa; regular at Kagoro forest in Nigeria (Dyer et al. 1986) and occur at density of one singing bird/12 ha in forest in Gabon (Brosset and Erard 1986). Most records are in the Congo basin.

Habitat and general habits Primary forest canopy, old secondary and gallery forest; in lowlands (Chapin 1939). In southwestern Uganda in Semliki and Maramagambo forests they

are solitary and uncommon in thick primary forest and dense secondary stands from 700 to 1200 m (Stevenson and Fanshawe 2002).They usually remain high in forest canopy, where males can be identified by their song.

Food Insects, mainly caterpillars especially in large masses, and beetles; fruit (Chapin 1939, Friedmann 1968, Brosset and Erard 1986, Irwin 1988,AMNH, LACM).

Breeding In Liberia on 13 March 1990 a female had an oviduct egg, broken, pale green with blackish blotches (Dickerman 1994;AMNH 827449). Breeding behaviors are unknown, the cuckoos are presumably brood-parasitic.

African Emerald Cuckoo Chrysococcyx cupreus (Shaw, 1792) Cuculus cupreus Shaw, 1792, Museum Leverianum, p. 157. [Gambia] Other common name: Emerald Cuckoo. Monotypic.

Description ADULT: Male, upperparts and head to breast glossy green, wing black with green outer vane, inner secondaries green, tail graduated in shape, the long inner T1 green,T2 green,T3 and T4 green with white tip, T5 green with broad white tip and white bars; under-

parts, lower breast and belly bright yellow, wing below with broad white patch formed by trailing edge of primaries and secondaries, under wing coverts bright yellow, base of flight feathers white and tip dark from below; eye-ring blue, iris brown, bill green above and slate gray below, feet bluish. Female, above head brown with whitish or dark brown flecks and bars, back to rump and upper tail coverts barred glossy green and rufous brown, wing coverts barred green and rufous, wing brown with rufous notches on outer vane and white bars on inner vane, tail coppery brown and white,T1 and T2 unmarked coppery

African Emerald Cuckoo Chrysococcyx cupreus 401 brown,T3 with rufous bars and a white tip,T4 white with 1–2 black bars and spots on inner and outer vane, outer T5 white, sometimes with a narrow black bar near the base and near the tip, face whitish finely barred brownish; underparts white with narrow green bars, breast washed buff, wing from below brown with narrow white bars across the flight feathers, under wing coverts white with narrow green bars; eye-ring bluish green, iris dark brown, bill black with edge of gape pale bluish green, feet bluish. JUVENILE: Sexes alike, head barred green and white, back to rump barred bright green and rufous brown, wing coverts and wing barred green and rufous brown, the green on inner vanes of back and wing feathers by dark brown, tail with outer tail feathers unbarred white or with 1–2 black spots or bars; underparts barred white and green (green bars as wide as white), underside of wing and wing coverts white barred greenish black; iris dark brown, bill black, feet gray. NESTLING: Naked at hatching, skin pinkish yellow then darkens with age; upper mandible whitish, gape orange. SOURCES: AMNH, BMNH, BWYO, CM, FMNH, LSU, NMGL, ROM, SMF, SMTD, UMMZ, USNM.

Geographic variation Subspecies C. c. intermedius Hartlaub 1857, C. c. sharpei van Someren 1922 and C. c. insularam Moreau and Chapin 1951 have been recognized based on tail length and by markings on the under tail coverts (less barred with white in the Gulf of Guinea islands), but these traits overlap between regions, and no subspecies are recognized.

Measurements and weights NE Zaire, Uganda and Kenya: Wing, M (n ⫽ 8) 98–108 (103.6 ⫾ 3.5), F (n ⫽ 8) 97–110 (101.9 ⫾ 4.2); tail, M 84–106 (95.1 ⫾ 7.1), F 71–91 (82.4 ⫾ 6.5); bill, M 15.2–17.2 (16.5 ⫾ 0.8), F 15.6–17.6 (16.7 ⫾ 0.7); tarsus, M 13.8–16.3 (15.2 ⫾ 0.9), F 14.7–16.1 (15.5 ⫾ 0.6) (AMNH); Cameroon: Wing, M (n ⫽ 10) 102–110 (105.9 ⫾ 2.6), F (n ⫽ 7) 99–13 (100.6 ⫾ 1.7); tail, M

93–109 (99.1 ⫾ 5.5), F 73–84 (77.7) (BMNH); Liberia: Wing, M (n ⫽ 8) 98–111.5 (105.5 ⫾ 4.8), F (n ⫽ 8) 97–101.5 (99.6 ⫾ 1.8); tail, M 100–116 (109.8 ⫾ 5.1), F 80–100 (85.9 ⫾ 7.3) (BMNH); Gulf of Guinea islands, Bioko: Wing, M (n ⫽ 9) 102–111 (106.8 ⫾ 2.6), F (n ⫽ 4) 99–102 (100.0); tail, M 92–102 (95.9 ⫾ 3.2), F 76–80 (77.5); Príncipe: Wing, M (n ⫽ 8) 104–111 (107.1 ⫾ 2.0), tail, 89–101 (93.5 ⫾ 3.7); São Tomé: Wing, M (n ⫽ 12) 104–109 (106.5 ⫾ 1.9), tail, 88–103 (94.7 ⫾ 4.1) (AMNH, BMNH, SMTD); Ghana: Wing, M (n ⫽ 14) 102–115 (108.4 ⫾ 4.0), F (n ⫽ 4) 104–107 (104.6); tail, M 92–121 (112.0 ⫾ 10.5), F 75–90 (82.3) (AMNH, BMNH, MSNM); Ethiopia: Wing, M (n ⫽ 20) 103–115 (108.6 ⫾ 3.1), F (n ⫽ 3) 99–107 (103.3); tail, M 92–133 (112.6 ⫾ 10.5), F 81–94 (88.3) (AMNH, BMNH, SMTD, USNM); South Africa: Wing, M (n ⫽ 15) 105–117 (109.1 ⫾ 4.1), F (n ⫽ 1) 104; tail, M 82–107 (93.0 ⫾ 6.0), F 82 (AMNH, BMNH, UMMZ). Weight, M (n ⫽ 21) 33–46 (38.3), F (n ⫽ 13) 30–46.3 (37.4) (UMMZ, Irwin 1988). Wing formula, P9 ⱖ 8 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 1 0 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 20 cm. Male glossy green with yellow lower breast and belly, female barred green and rufous above, wing green barred rufous, outer rectrices white with black spots, tail coverts without wide white edge, below white with green bars. In flight, male with under wing coverts and broad band on base of flight feathers unmarked and pale, tips of flight feathers dark, female like female Klaas’s Cuckoo C. klaas. Juveniles are more strongly barred on head than female, and differ from juvenile C. klaas by crown barred green and white (green and rufous in C. klaas), broad green bars below (thin bronze bars in C. klaas), rump barred rufous and green (buff and dark brown in C. klaas), and fewer black bars on white outer tail feathers.

Voice Song, clear four-part bouncy, rhythmic slurred whistle, “hello, Ju-dy” or “goodby, Aunt Bet”, the

402 African Emerald Cuckoo Chrysococcyx cupreus second note low in pitch, the last note highest, the intonation as in British English (not U.S. English), consistent across species range in Africa (Chappuis 1974, 2000, Stjernstedt 1993, RBP). Other calls are long clear whistles and short clear whistled trills (Chappuis 2000).

Range and status Africa from The Gambia and southern Senegal to Ethiopia, Central and East Africa to South Africa; Gulf of Guinea islands (Chapin 1939, Friedmann 1968, Irwin 1988; Maclean 1993, Zimmerman et al. 1996). Seasonal migrant in southern Africa, some birds may remain in dry season when they are silent and difficult to observe (Rowan 1983,Vernon et al. 1997, Aspinwall and Beel 1998). Mainly resident in East Africa. Local resident or partial migrants with the rains in West Africa, resident in forest areas and arrive with rains in gallery forests in the northern savannas (Thiollay 1985, Grimes 1987, Cheke and Walsh 1996, Elgood et al. 1994), seen and heard all year at Kumba, Mt Kupé and Yaounde in Cameroon (Serle 1954, Quantrill and Quantrill 1998, Bowden 2001). Fairly common, high in dense foliage, conspicuous when they sing. In the 19th century these cuckoos were common in The Gambia, Casamance and Togo; now they are rare (Gore 1990, Morel and Morel 1990, Cheke and Walsh 1996).

Habitat and general habits Evergreen forest, seasonal miombo woodland, thickets, large shady trees, forest clearings and edges, cocoa plantations under shade trees, Cecropia trees, open woodlands, gallery forest, in towns and

suburbs where they sing in introduced eucalypts (Lilongwe, Malawi); occur in lowlands in West Africa, in forested belt at 1300–1500 m in Gulf of Guinea islands (Pérez de Val 1996, Christy and Clarke 1998), to 1000 m in Liberia (Gatter 1997), in Kenya from lowlands to 2000 m highland forest “islands” in semi-arid regions, most in 500⫹ mm rainfall region (Lewis and Pomeroy 1989). Outside rainy season it is in closed canopy and liana-clad forest. Usually remain high in trees, over 30 m in high forest, where they forage in dense canopy (Chapin 1939, Christy and Clarke 1994, Gatter 1997), alone, or in a group when feeding on concentration of caterpillars (Irwin 1988, Leonard 1998). Male sometimes holds a caterpillar in the bill as he sings.

Food Insects, mainly hairy caterpillars, and Acraea butterflies (Irwin 1988). Occasionally takes fruit and tree seeds (Rowan 1983, BMNH).

Displays and breeding behavior Male sings, female arrives and perches, male takes short flights from branch to branch, lands near her, raises and spreads its tail, droops its wings and moves sideways along the branch, singing, the female droops and quivers its wings, and the pair copulates. Male then flies with conspicuous wingbeats to a tree, picks off insects, flies to the female and feeds her (van Someren 1956). Male sometimes feeds a fledged young cuckoo (Worman 1930, Friedmann 1956, 1968). Female lays in sets of 3–4 eggs, on successive days or alternate days, ovulating in the morning, and lays about 20 eggs in a season (Payne 1973a, 1974).

Breeding In Liberia the males sing from March to October (Gatter 1997) and a female had an oviduct egg in August (Colston and Curry-Lindahl 1986); in Cameroon, at Kumba a female had large ovarian follicles in March (Serle 1954) and at Mt Kupé the species sings all year but more frequently from July to December (Bowden 2001), in São Tomé it lays early October-late December (de Naurois 1979), in Gabon they lay September to March (Brosset and

Long-billed Cuckoo Chrysococcyx megarhynchus 403 Erard 1986), in Kenya they lay during the long rains when the hosts breed (Brown and Britton 1980), in Angola a female with an oviduct egg was taken in February (Heinrich 1958) and in southern Africa they breed during the rains mainly in October and November (to January) ( Jensen and Jensen 1969, Benson et al. 1971, Benson and Benson 1977, Rowan 1983). Brood-parasitic. Hosts, mainly insectivorous small songbirds, including bulbuls, warblers, robin-chats, flycatchers, sunbirds and weavers (Friedmann 1948, 1968, Payne and Payne 1967). Host species are not well known, as cuckoos were often identified from eggs not seen to hatch and develop into young with a distinctive plumage. In South Africa, the best documented hosts are Starred Robin Pogonocichla stellata, Bleating Bush Warbler Camaroptera brachyura and Blue-gray Flycatcher Muscicapa caerulescens ( Jensen and Jensen 1969, Vernon 1970b, Oatley 1980, Rowan 1983, Maclean 1993). In Mozambique, young cuckoos are reared by Southern Puffback Shrike Dryoscopus cubla (Vincent 1934). In Gulf of Guinea, on Bioko the hosts are white-eyes Zosterops and Speirops (Keulemans 1907, in Friedmann 1956 and de Naurois 1979), on São Tomé the hosts are

Newton’s Sunbird Nectarinia newtonii and São Tomé Weaver Ploceus sanctithomae (de Naurois 1979, 1994). In Central Africa,Yellow-whiskered Greenbul Andropadus latirostris rear the young cuckoo on a diet of fruit late in the nestling stage but feed it insects during the first days and then feed a mix of pulped fruit and insects (van Someren 1956, Brosset 1976). Other hosts in Gabon and Congo (Brazzaville) are insectivores: illadopsis Illadopsis, paradise-flycatchers Terpsiphone, wattle-eye flycatchers Platysteira and sunbirds Nectarinia (Vincent 1965, Brosset 1976, Brosset and Erard 1986). Eggs are 20 ⫻ 15 mm (Irwin 1988, Birds of Africa; cf. vol. 5, 1997, p. 643), 22 ⫻ 14 mm, weight 2.1 g (UMMZ 216725). Eggs are white, pink or pale blue, usually marked with brown, or bluish gray with brown spots and blotches, or rose red with brown cap, or rose salmon with circle of red speckles, with limited evidence of mimicry of host eggs (Tarboton 2001).The female cuckoo removes a host egg from the nest. Incubation period is 13–14 days. The nestling cuckoo evicts the host eggs and young from the nest.The nestling period is 18–20 days, and the young cuckoo is in the care of its foster parents for 2 weeks after fledging (Brosset 1976, Oatley 1980).

Long-billed Cuckoo Chrysococcyx megarhynchus (G. R. Gray, 1858) Cuculus megarhynchus G. R. Gray, 1858, Proceedings of the Zoological Society of London, 1858, p. 184. (Aru Islands) Monotypic. Other names: Rhamphomantis megarhynchus (G. R. Gray, 1858).

Description ADULT: Sexes unlike, male above head blackish, back dark reddish sooty brown, wing black, tail black, cheek to malar region black; underparts, throat brownish, breast blackish, belly to under tail coverts rufous brown, underside of tail brown with indistinct blackish bars; eye-ring red, iris red, long black narrow bill decurved at the tip, feet grayish blue. Female above, head dark gray-brown, back dull cinnamon brown, cheek to malar region dark

gray; underparts, throat and breast gray finely barred buff, lower breast rufous, belly and under tail coverts either rufous or light gray finely barred and vermiculated rufous; eye-ring inconspicuous gray, iris dark brown with a narrow cream outer ring. JUVENILE: Cinnamon head and back, face gray and brown, blackish ear coverts, pale gray around the eye and behind the ear; underparts brown, the throat pale, breast sooty gray, belly creamy brown, unbarred and without vermiculation, tail with an incomplete blackish subterminal bar; eye-ring dark, iris brown, bill black to brown, gape light yellow, feet grayish brown. NESTLING: Undescribed. SOURCES: AMNH, BMNH, MZB, RMNH, SMTD, USNM,YPM, ZMB.

404 Long-billed Cuckoo Chrysococcyx megarhynchus

Measurements and weights New Guinea and Aru Islands: Wing, M (n ⫽ 4) 96–100 (97.8), F (n ⫽ 6) 96–104 (99.1 ⫾ 2.0); tail, M 72–84 (79.1), F 66–82 (76.2 ⫾ 4.8); bill, M 22.4–24.8 (23.6), F 22.2–25.5 (22.6); tarsus, M 15.8–16.2 (16.0), F 14.8–18.6 (16.3) (AMNH, BMNH, MZB, RMNH, SMTD, USNM,YPM, ZMB). Weight, F (n ⫽ 1) 31.5 (USNM), U juvenile (n ⫽ 1) 24 (Ripley 1964,YPM). Wing formula, P8 ⫽ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10. Geographic variation: A subspecies C. m. sanfordi (Stresemann and Paludan, 1932) was described from Waigeo Island, on the basis of a female with a gray breast contrasting with a rufous belly; males from Waigeo are undescribed (Rothschild et al. 1932).The holotype of sanfordi (AMNH 300703) is similar to a breeding female (USNM 584998) from Korhoro, 50 km SSW of Wau, Papua New Guinea, so the birds of Waigeo are not distinct and the subspecies is not recognizable. Birds on the Vogelkop Peninsula and Misool may differ (Gyldenstolpe 1955, Rand and Gilliard 1967), although the few specimens available do not support the existence of a subspecies.

Field characters Overall length 18 cm. Small cuckoo, it perches in an upright posture, the bill long with a drooping tip, tail short for a cuckoo. Inconspicuous, elusive and solitary. Flight is straight and direct, like other bronze-cuckoos (Ripley 1964). When the bird is feeding or moving about the body is held horizontally like that of a cuckoo; when it is perched the vertical posture and long bill suggest a honeyeater, and the immature plumage is marked like a Tawnybreasted Honeyeater Xanthotis flaviventer. The adult cuckoo is shaped, colored and marked like a female Twelve-wired Bird of Paradise Seleucidis melanoleuca, but is much smaller.

Voice A loud, distinct trill, a descending series of evenly spaced sibilant notes, the downslur lasting 4 sec, repeated at 5-sec intervals, soft but carries far, suggests trilled song of Little Bronze-cuckoo Chrysococcyx minutillus; and a sweet two-note whistle, the

first note c. 1 sec, second note c. 2 sec and upslurred (Coates 1985, 1991, Beehler et al. 1986, J. Diamond); also described as an accented three-note whistle “hüo HÜTT tühü” (Stein 1936).

Range and status New Guinea, the western Papuan islands of Misool and Waigeo, and Aru Islands. In New Guinea they occur in the lowlands at Vogelkop, Mamberamo River and Humboldt Bay in Irian Jaya; and the Sepik River, Upper Fly and Kumusi River and Port Moresby region in Papua New Guinea. Resident, perhaps nomadic. This is the rarest cuckoo on New Guinea (Gyldenstolpe 1955,Ripley 1964,Coates 1985,1991).

Habitat and general habits High 70 m forests and openings in lowland rainforest, monsoon forest, forest edge and secondary growth of lowland New Guinea. Forage in canopy and sub-canopy, motionless for long periods, they glean foliage, branches and bark, and appear at outbreaks of caterpillars where they associate with other birds (Stein 1936, Rand and Gilliard 1967, Coates 1985, Beehler et al. 1986).This species is one of the least known and rarest birds in New Guinea.

Food Insects, mainly caterpillars, and flying insects, perhaps ants (Coates 1985).

Breeding A female (USNM 584998) was taken in September with an oviduct egg, undescribed, ready to lay. Presumably brood-parasitic.

Horsfield’s Bronze-cuckoo Chrysococcyx basalis 405

Horsfield’s Bronze-cuckoo Chrysococcyx basalis (Horsfield, 1821) Cuculus basalis Horsfield, 1821, Transactions of the Linnean Society of London, 13 (1), p. 179. ( Java) Monotypic. Other common names: Australian Bronzecuckoo, Narrow-billed Bronze-cuckoo, Rufoustailed Bronze-cuckoo. Other names: Chalcites basalis (Horsfield).

Description ADULT: Sexes alike, above the head to back and rump brown, sometimes with bronze wash on back, wing coverts gray with bronze green gloss and buff edges, wing gray brown with buff edges in fresh plumage, tail blackish with rufous, the inner feathers T1 unspotted gray glossed green,T2 to T4 rufous at base and center and rufous edge, T2 and T3 gray with subterminal blackish band and broad white tip, T4 with two white spots, T5 black with five broad white bars on inner vane, the outer vane gray with indistinct marks, face whitish, dark eye streak with white eyebrow and whitish cheek, thin black malar streak; underparts whitish, throat lightly streaked, sides of breast and flanks barred light bronze, the bars incomplete in middle of breast and belly, underside of flight feathers with pale rufous white trailing vane, under wing coverts white with light bronze bars; eye-ring gray, iris red (males) to brown (males and females), bill slender and black, feet gray. JUVENILE: Head and back gray brown, wing gray brown, tail dark gray with rufous base and edge of tail feathers, face gray, dark gray below the eye and a slightly darker line through the eye; underparts, chin to breast light gray, the breast, belly and under tail coverts whitish, flanks faintly barred; eye-ring gray, iris gray to light brown. NESTLING: Naked at hatching, skin pink and gray, darkens with age; gape flanges white, mouth lining yellowish (Brooker and Brooker 1986, Payne and Payne 1998a). SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, MZB, ROM, UMMZ, USNM,WAM, ZRC.

Measurements and weights Wing, M (n ⫽ 78) 97–108 (102.7 ⫾ 2.4), F (n ⫽ 32) 95–108 (100.7 ⫾ 2.8); tail, M 64–75 (70.9 ⫾ 2.2), F 63–76 (69.3 ⫾ 2.8); bill, M 14.0–18.3 (16.7 ⫾ 0.9), F 15.2–18.7 (17.2 ⫾ 0.9); tarsus, M 16.2–19.8 (18.0 ⫾ 0.9), F 15.9–19.2 (17.9 ⫾ 0.7) (Higgins 1999);Western Australia: Wing, M (n ⫽ 15) 96–106 (100.2 ⫾ 3.0), F (n ⫽ 7) 94–103 (99.1 ⫾ 2.7); tail, M 68–75 (71.4 ⫾ 2.6), F 67–76 (70.3 ⫾ 3.2) (WAM, FMNH). Weight, M (n ⫽ 22) 17.0–30.0 (22.0), F (n ⫽ 16) 20.0–32.0 (24.3) (Higgins 1999). Wing formula, P9 ⬎ 8 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 1 0 ⬎ 2 ⬎ 1. The two common bronze-cuckoos in Australia were recognized as distinct species by Gould (1845b), who emphasized the narrow bill shape and the rufous tail in C. basalis.

Field characters Overall length 17 cm. Small cuckoo with bronze upperparts and a gray streak behind the eye, rufous edge to base of tail, the tail partly rufous below, the underparts variably barred. Differs from Shining Bronze-cuckoo C. lucidus in face with a dark eye streak, less brightly colored upperparts, rufous base of outer tail T2 to T4,T4 unbarred except near tip, and underparts less barred. Juvenile differs from Shining Bronze-cuckoo C. lucidus in nearly uniform gray head and throat with a dark streak through the ear (Australian C. lucidus have a whiter face, lack the streak and are speckled on the throat), rufous base of tail and the underparts only indistinctly barred. Distinguished by its high-pitched, downward inflected whistles (the whistles rise in C. lucidus), the call faster and higher than that of Black-eared Cuckoo C. osculans.

Voice Descending high whistles “peeer”, plaintive and penetrating, each note dropping from 4.2 to 2.2 kHz, the whistle incessantly repeated about once in a second. Another call, a “chirrup” (Buckingham and Jackson 1990). Begging calls of nestling and fledgling cuckoos

406 Horsfield’s Bronze-cuckoo Chrysococcyx basalis differ with the host and the variation suggests genetic races specialized towards the major host groups. In Western Australia the begging calls of young cuckoos reared by Splendid Fairy-wrens Malurus splendens are at first a “peep” then change to “ze-ba” and “sit” calls and a “twe” that falls then rises by late nestling life; after fledging the birds combine “twe” calls into a reel like a young fairy-wren. In contrast, begging calls of fledged cuckoos reared by Western Thornbills Acanthiza inornata give a whine like a young thornbill. The begging calls of young reared by less regular hosts such as the irruptive Australian robins Petroica which appear in the season after a grass fire, match one of these two forms (Payne and Payne 1998a, Figure 10.5). In northern Queensland a fledgling with Red-backed Fairywrens Malurus melanocephalus hosts and in New South Wales a fledgling with Superb Fairy-wren M. cyaneus hosts, each had a call like those of young fairy-wrens (Goodwin 1974, Alley 1978). In southeastern Australia (Buckingham and Jackson 1991), calls of fledgling cuckoos in Victoria begging from Brown Thornbill Acanthiza pusilla are a thornbill-like buzz, like calls of fledglings with Western Thornbills in Western Australia; the young cuckoos reared by Superb Fairy-wrens Malurus cyaneus sound like the young cuckoos with Splendid Fairy-wrens in Western Australia.

Range and status Australia and Tasmania. Partial migrant. A few birds remain in winter in Australia east of the Dividing Range and in the south; more remain in winter in northern Australia. In Tasmania occur from June to March, in SE and SW Australia from June to April (Pizzey and Knight 1998, Higgins 1999). Birds occur on islands off the coast in migration, and winter north of Australia. In austral winter from March to October, cuckoos occur in New Guinea (rare in SE New Guinea), Christmas I, Aru Islands; in Wallacea rare in Sulawesi (Palu Valley, Lompobattang Massif, and south coast of northern peninsula), Lesser Sundas (Flores where common April and May, Lombok, Sumbawa, Kalaota, Timor), Greater Sundas (three records in Sumatra, many in Java, once in Borneo), the Moluccas (Halmahera), Christmas I, Natuna I and Singapore (Changi coast) (Smythies 1957, 1981, Coates 1985,

White and Bruce 1986, van Marle and Voous 1988, Johnstone et al. 1993, Coates and Bishop 1997,Verhoeye and Holmes 1998,Wells 1999, Higgins 1999). Common in much of Australia, scarce in Northern Territory and Kimberley Division (Rensch 1931, Coates 1985, Higgins 1999). Regions with no recent observations have host populations that breed in years of good conditions (e.g. Rowley and Russell 1997), and cuckoos may then occur. Breeding densities, estimates of 0.1 bird/ha, or 10/km2 (Blakers et al. 1984) are high, more often c. 2/km2 (Brooker and Brooker 1992, 1994; color-marked birds at Gooseberry Hill, Western Australia, Payne and Payne 1998a).

Habitat and general habits Open woodland, mallee, mulga, scrub, dry interior spinifex, coastal saltmarsh, saltbush, generally arid and semiarid zones (Pizzey 1980, Brooker and Brooker 1992, 1994); in winter occurs in scrub and secondary woodland, and dry open woodland, mainly coastal and subcoastal zones from sea level, occasionally in the mountains to 1830 m (Sulawesi) (Coates and Bishop 1997). Usually feed singly, occasionally in flocks in winter. Forages in tree and bush foliage and on the ground (Bravery 1970, Pizzey 1980, Sody 1989, Higgins 1999).

Food Insects, mainly caterpillars, especially ground caterpillars, and beetles (Coccinellidae), bugs; eggs of small birds (Higgins 1999).Young cuckoos are fed by host parents on insects, including long-horned grasshoppers and caterpillars (Alley 1999).

Black-eared Cuckoo Chrysococcyx osculans 407

Displays and breeding behavior Male is territorial and sings in an area about 1 km2. There is no apparent interaction between this cuckoo and other glossy cuckoos that breed in its habitat (C. lucidus, C. osculans). The male feeds the female in courtship, and occasionally he feeds a fledged young cuckoo (Friedmann 1968, Brooker and Brooker 1989,Tarburton 1993).

Breeding In eastern Australia they lay from July to January, in South Australia from July to February. In Western Australia the season varies with the regional rains, in southwestern Division from August to October, near Perth from October to December, in Kimberley Division from January to March, and in Pilbara Region from March to May and August to September, while in the arid inland the cuckoos breed in spring and in autumn (Storr 1984a, Brooker and Brooker 1989a,b, Ambrose and Murphy 1994, Pizzey and Knight 1998, Higgins 1999). Brood-parasitic.The hosts are mainly small thornbills Acanthiza spp., fairy-wrens Malurus spp., Australian robins Petroica spp. and Australian chats Ephthianura spp.; 28 species are known to have reared a young cuckoo, mainly species that lay in covered nests (76% of records), and some hosts that lay in open nests (24%) (McGilp 1941, Storr 1977, 1980,

1984a,b, 1991, Brooker and Brooker 1989b, Strahan 1994). Eggs are whitish with brown flecks, 18 ⫻ 12 mm (Serventy and Whittell 1976, Beruldsen 1980). Laying takes only 6 sec, when the female cuckoo enters the nest part way, lays her egg, then backs out with a host egg in the bill (Brooker et al. 1988). Cuckoo eggs that are laid before the host eggs, are sometimes buried in the nest lining as the host continues to build (McGilp 1941). Incubation period is 11–13 days. The nestling evicts host eggs and nestlings usually at 2 days after hatching (it can evict at 4 days, it does not at 5 days).The pinfeathers emerge on day 6, the tail feathers emerge from the sheaths on day 9 and the primaries on day 10, the chick is well feathered by day 14 and it fledges at 20–21 g on day 17–19. The young are fed by their foster parents for up to four weeks after fledging, although they begin to feed themselves in three days after fledging (Barrett 1905,Alley 1978, Brooker and Brooker 1989b, Payne and Payne 1998a, Higgins 1999). Success, of 95 cuckoo eggs laid in nests of Splendid Fairy-wren Malurus splendens, 71 (74.7%) hatched and 41 (43%) fledged; half the failures were due to nest predation, while the others failed due to burial of the egg in the nest lining, late laying, accidents, infertility and desertion. In contrast to the natural situation, in an experimental study of superb fairy-wrens M. cyaneus, the nesting birds often deserted the cuckoo chicks (Langmore et al. 2003).

Black-eared Cuckoo Chrysococcyx osculans (Gould, 1847) Chalcites osculans, Gould, 1847, Proceedings of the Zoological Society of London, 1847, p. 32. (New South Wales) Monotypic. Other names: Misocalius osculans (Gould).

Description ADULT: Sexes alike, above uniformly gray with slight metallic gloss, wing coverts gray, wing gray, rump pale gray, tail blackish with narrow white tips and rufous cast in fresh plumage, face with white streak above the eye, broad black streak through the eye; underparts white with rufous tinge in fresh plumage, the under wing coverts pale and unmarked,

under wing gray, inner webs of inner flight feathers and coverts white barred gray-brown and inner secondaries unbarred pale gray, form a pale patch when wing is seen from below in flight; eye-ring gray, iris black, bill black, feet dark gray. JUVENILE: Similar to adult, slightly duller, face marks less distinct than in adult, face with a brown smudge in front of the eye and a dark brown patch on the ear coverts, lower belly and under tail coverts buff; bill with pink palate and gape flanges yellow. NESTLING: Naked at hatching, skin black.

408 Black-eared Cuckoo Chrysococcyx osculans SOURCES: AMNH, BMNH, FMNH, MVZ, UMMZ,WAM.

Measurements and weights Wing, M (n ⫽ 26) 114–123 (118.4 ⫾ 1.9), F (n ⫽ 16) 113–122 (116.9 ⫾ 2.3); tail, M 83–92 (88.2 ⫾ 2.4), F 83–92 (87.3 ⫾ 2.3); bill, M (17.3– 21.3 (19.5 ⫾ 1.1), F 17.1–21.2 (19.5 ⫾ 1.1); tarsus, M 16.9–20.9 (19.2 ⫾ 0.8), F 18.1–22.0 (19.4 ⫾ 0.9) (Higgins 1999);Western Australia:Wing, M (n ⫽ 5) 112–118 (115.8 ⫾ 2.4), F (n ⫽ 8) 112–117 (114.1 ⫾ 1.8) (WAM). Weight, M (n ⫽ 9) 25.5–35.4 (29.4 ⫾ 3.5); F (n ⫽ 5) 27–38.5 (32.2 ⫾ 5.0) (Higgins 1999). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 20 cm. Small cuckoo, gray above and white below, with white streak above the eye and a broad black streak through the eye. Differs from other small cuckoos in Australia by lack of metallic plumage above, rump gray not green, tail gray not rufous, and lack of barring on underparts and under wing coverts. C. osculans differs from juvenile Horsfield’s Bronze-cuckoo C. basalis in paler plumage, lacks scaly upper wing coverts, and has rump and upper tail coverts paler than back. Both C. osculans and C. lucidus lay unmarked brownish eggs; the eggs of C. osculans are dark brown and those of C. lucidus are bronze.

Voice Song, a descending low whistle “peeeer” (lower in pitch than Horsfield’s Bronze-cuckoo C. basalis). Courtship call, a repeated “peeowit-peeoweer”, and calls in courtship behavior (Buckingham and Jackson 1990, Higgins 1999). Fledged young give a “chirp”.

Range and status Australia. Migratory in most of the breeding range in Australia, or nomadic with movements varying with irregular rains, breed with the host species in response to rains. Breeds mainly south of 23°S in Australia. Breeding visitor from September to April in northern Queensland, from April to June in southern coastal plains in Queensland, wanders to northern Australia, and some remain in winter in

southern Australia (Storr 1980, 1984b, Blakers et al. 1984, Higgins 1999); vagrant in Tasmania.Vagrant or rare wintering migrant in southwestern New Guinea,Aru Islands, the Moluccas and eastern Lesser Sundas (Bacan, Kai Islands, Luang, Babar, Romang in July, September, November) (van Bemmel 1948, Beehler et al. 1986, White and Bruce 1986, Coates and Bishop 1997). Fairly common, except in wet, forested coastal areas where they are uncommon.

Habitat and general habits Open scrub, dry woodlands, mallee, mulga, saltflats, waterside forests and thickets, mainly in low shrubby vegetation; rare in subhumid zone. Perch in low trees and shrubs, forage in shrubs and on ground. Flight is swift and direct.

Food Insects, mainly caterpillars, and beetles, diptera, sandflies and hemiptera; seeds ( Johnstone and Storr 1998, Higgins 1999).

Displays and breeding behavior Female calls, male approaches, female calls shorten and increase in tempo, male calls, female quivers her wings and crouches on a branch, then both male and female call with a higher pitch, and copulate. The male feeds the female in courtship (Rix 1976, Higgins 1999).

Breeding Season varies with regional rains in Australia, in the west from June to October and in the east from August to December (Storr 1984a, Higgins 1999).

Rufous-throated Bronze-cuckoo Chrysococcyx ruficollis 409 Brood-parasitic. Hosts, small Australian warblers that nest on or near the ground (⬍ 0.5 m), mainly Redthroat Sericornis brunneus and Little Fieldwren S. sagittatus, and other species with covered nests, some of these are hosts with spotted eggs unlike its own. Female removes a host egg when she lays her

own. Eggs are dark brown, 20 ⫻ 14 mm, and closely resemble Redthroat eggs. Incubation period is unknown. Nestling period is 18 days. The nestling evicts the host eggs and young (Chisholm 1973, Beruldsen 1980, Brooker and Brooker 1989a, Higgins 1999).

Rufous-throated Bronze-cuckoo Chrysococcyx ruficollis (Salvadori, 1875) Lamprococcyx ruficollis Salvadori, 1875, Annali del Museo Civico di Storia Naturale di Genova, 7, p. 913. (Hatam, Arfak Mountain [New Guinea]) Other common names: Reddish-throated Bronze Cuckoo, Mountain Bronze Cuckoo. Other names: Chalcites ruficollis (Salvadori). Monotypic.

Description ADULT: Sexes alike, head and back coppery green, wing coverts coppery green, wing greenish with rufous inner vane of flight feathers, tail bronze above on T1 and T2;T3 to T4 barred rufous on the inner vane, T3 to T5 have white tip, T5 is barred white on inner and outer vane, forehead and face light rufous; underparts, throat and breast light rufous, chin, throat and breast to belly and under tail coverts barred bronze green, the dark bars 2–3 mm wide; bend of wing white, under wing coverts white with dark bars like underparts, a white wing bar across the inner vane of P1–P7 and the secondaries, P8 all dark, P9 with white at base of inner vane; eyering gray-green, iris reddish brown (male) to brown (female), bill black, feet olive gray. JUVENILE: Above greenish; underparts, the throat and breast light gray, belly white, flanks faintly barred gray; iris dark brown. NESTLING: Undescribed. SOURCES: AMNH, KU, MCZ, MSNG, RMNH, YPM, ZMB.

Measurements and weights Wing, M (n ⫽ 9) 89–101 (94.8 ⫾ 3.6), F (n ⫽ 9) 94–96 (95.5 ⫾ 1.2); tail, M 58–72 (65.0 ⫾ 4.6), F

60–69 (63.8 ⫾ 2.9); bill, M 13.1–15.0 (14.1 ⫾ 0.7), F 13.2–15.8 (14.0 ⫾ 0.8); tarsus, M 15.0–16.4 (16.0 ⫾ 0.5), F 14.4–16.7 (15.7 ⫾ 0.8) (AMNH, BMNH, KU, MCZ, MSNG, RMNH,YPM, ZMB). Weight, M (n ⫽ 1) 21.0, F (n ⫽ 1) 23.5 (laying) (AMNH,YPM). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⱖ 10 ⬎ 2 ⬎ 1. A specimen label note on a breeding female collected by Mayr in 1928, AMNH 293683, is “L.R. graugrun”, where L.R.” refers to “Lid Rand” or eye-ring (E. Mayr, in litt.). Hartert (1930) listed the specimen but did not note the detail; the eye-ring color was not previously described (Parker 1981).

Field characters Overall length 16 cm. Upperparts glossy green, underparts white barred rufous, the bars wider than in C. minutillus, face and throat rufous, without white (forehead and face with white in C. minutillus, chin and throat white in Little Bronze-cuckoo C. minutillus and White-eared Bronze-cuckoo C. meyeri), eye-ring gray-green (red in C. minutillus). Juvenile has green upperparts (not brown as in C. meyeri) and is only slightly rufous below.

Voice Song, 8–9 identical high-pitched, downslurred whistles, 2/sec, higher-pitched than White-eared Bronze-cuckoo C. meyeri. Also a single falling note “tseew” (Beehler et al. 1986).

Range and status New Guinea (Rothschild and Hartert 1907, Hartert et al. 1936, Parker 1981, Coates 1985, Beehler et al. 1986). Resident, uncommon to rare.

410 Shining Bronze-cuckoo Chrysococcyx lucidus elevations than Little Bronze-cuckoo C. minutillus and White-eared Bronze-cuckoo C. meyeri (Diamond 1972). They join mixed-species flocks of feeding birds. In foraging they perch motionless, staring and peering about for insects under small branches, then move and seize the insect, and move slowly from tree to tree (Mayr and Rand 1937), and they sally out from a high perch to catch insects in flight (Coates 1985).

Food Insects, mainly caterpillars, and flying insects (Coates 1985).

Habitat and general habits Montane forest and forest edge. Highlands of New Guinea, mainly 1800–2600 m, to 3350 m, at higher

Breeding A female had a large ovary in July (AMNH 293683). Presumably brood-parasitic, the host and eggs are unknown.

Shining Bronze-cuckoo Chrysococcyx lucidus (Gmelin, 1788) Cuculus lucidus Gmelin, 1788, Systema Naturae, 1, part 1, p. 421. (New Zealand) Other common names: Golden Bronze-cuckoo, Shining Cuckoo. Other names: Chalcites lucidus (Gmelin). Polytypic. Four subspecies. Chrysococcyx lucidus lucidus (Gmelin, 1788); Chrysococcyx lucidus plagosus (Latham, 1801); Chrysococcyx lucidus layardi Mathews, 1912; Chrysococcyx lucidus harterti Mayr, 1932.

Description ADULT: Sexes similar, differ when subspecies are taken into account. Male, C. l. lucidus, crown to back and rump glossy green, wing coverts glossy green, wing brown with inner secondaries glossed green, tail bronze green above with black subterminal band, T1 dark, T2 to T5 with white tip, T5 with inner vane barred black and large white spots, face pale with fine white spots on cheek; underparts, chin barred or mottled, throat white or narrowly barred greenish, breast and belly to under tail coverts white barred greenish to bronze, under wing coverts white with bronze bars. Female, cap brown to bronze, back green with bronze gloss, ear

coverts duskier, face with less white than male, chin pale with little or no barring; underparts barred as in male; narrow eye-ring gray, iris gray to dark yellow to brown (some with gray outer and red inner ring). In Western Australia the iris white or pearl gray in male, brown in female, bill black (pale base in female), feet dark gray. JUVENILE: Above dull green, wing brown with buff edges, tail dark gray with no rufous, face pale, whitish above and below the eye; underparts, throat whitish, breast light gray, belly white, flanks and sometimes belly inconspicuously barred; eye-ring gray, iris gray to pale brown, bill black, feet dark gray to olive brown to dark brown. NESTLING: New Zealand and New Caledonia, at hatching head and back sparsely covered with white hair-like down; Australia, at hatching naked or nearly naked; skin pinkish orange to greenish gray; gape flange white in New Zealand, yellow in SW Australia. SOURCES: AMNH, BMNH, CM, FMNH, MVZ, ROM, UMMZ, USNM,WAM, ZFMK, ZMUC.

Shining Bronze-cuckoo Chrysococcyx lucidus 411

Subspecies Chrysococcyx lucidus lucidus (Gmelin, 1788); male, crown and upperparts uniform green, extensive white ear coverts, forehead, sides of crown and face with much white spotting; female, crown bronze, back green with bronze gloss, ear coverts duskier, less white on face than male, chin pale with little or no barring, bill broad and deep; New Zealand and Norfolk I; migrates to Solomon Islands, Bisrnarck Archipelago and Woodlark I; Chrysococcyx lucidus plagosus (Latham, 1801); male plumage like female plumage of New Zealand birds, head more brown than the back, slightly less green on upperparts than New Zealand birds. bill narrower, female lacks white spots on forehead and over the eye, crown bronze, upperparts bronze, side of face less white, without white scallops on the forehead and with more brown or bronze (less green) gloss on back and on ventral bars, bill narrow and shallow; Australia and Tasmania, migrates to winter in New Guinea and Lesser Sundas; Chrysococcyx lucidus layardi Mathews, 1912; smaller, crown brown without white spots, face dark (adult and juvenile), dark extends below the eye, throat white and unbarred, the breast and belly more finely barred than in C. l. lucidus; bill broad; New Caledonia,Vanuatu ( ⫽ New Hebrides), Banks I, Santa Cruz Islands, Loyalty Islands. Birds on Malekula I, Vanuatu, described as Chrysococcyx lucidus aeneus (Warner 195l), do not differ from other birds of the island region; Chrysococcyx lucidus harterti Mayr, 1932; smaller, male crown copper and female crown purple, male back green and female back bronze, the throat unbarred, breast and belly more finely barred than in C. l. lucidus, male underparts white and female throat and upper breast dark rufous; Rennell and Bellona Islands.

Measurements and weights C. l. lucidus; New Zealand: Wing, M (n ⫽ 40) 101–112 (104.1 ⫾ 2.1), F (n ⫽ 38) 99–109 (104.2 ⫾ 2.0); tail, M 63–75 (68.7 ⫾ 2.3), F 64–73 (68.8 ⫾ 2.2); bill, M 16.9–19.7 (18.2 ⫾ 0.6), F 17.0–19.5 (18.6 ⫾ 0.6); tarsus, M 16.8–20.7 (18.6 ⫾ 0.7), F 18.0–19.1 (18.8 ⫾ 0.6) (Higgins 1999);

C. l. plagosus; Western Australia: Wing, M (n ⫽ 26) 99–110 (106.3 ⫾ 2.2), F (n ⫽ 11) 101–110 (105.6 ⫾ 2.9); tail M (n ⫽ 13) 65–72 (68.5 ⫾ 2.3), F (n ⫽ 4) 66–72 (69.3 ⫾ 2.5); bill M (n ⫽ 24) 16.5–18.9 (17.6 ⫾ 0.7), F (n ⫽ 8) 15.3–19.1 (17.3 ⫾ 1.3); tarsus M (n ⫽ 10) 16.0–18.0 (16.9 ⫾ 0.6), F (n ⫽ 3) 16.8–17.9 (17.4) (Higgins 1999); wing, M (n ⫽ 12) 100–109 (105.0 ⫾ 2.5), F (n ⫽ 8) 98–108 (104.3 ⫾ 3.2) (WAM); C. l. layardi: Wing, M (n ⫽ 17) 94–101 (99.5), F (n ⫽ 7) 93–100 (97.1); tail, M 61–73 (68.3), F 62–71 (67.2) (Mayr 1932, FMNH, UMMZ); C. l. harterti: Wing, M (n ⫽ 3) 90–93 (92.7), F (n ⫽ 3) 90–95 (93.0); tail, M 62–64 (63.0), F 60–63 (61.7) (Mayr 1932, ZMUC). Weight, C. l. lucidus: M (n ⫽ 16) 16.7–32.0 (25.6), F (n ⫽ 12) 18.5–29.0 (24.4); C. l. plagosus: M (n ⫽ 38) 17–35 (23.3), to 50 before migration, F (n ⫽ 30) 18.0–28.5 (24.5); C. l. layardi: M (n ⫽ 3) 21–21.5 (21.2) (Higgins 1999, UMMZ, ZSMK). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⱖ 1 0 ⬎ 2 ⬎ 1.

Field characters Overall length 17 cm. Small cuckoo with glossy green upperparts and a pale face, the face indistinctly freckled dark and the forehead sometimes flecked white, no rufous in the tail, and the underparts distinctly barred. In Australia, juvenile lacks rufous in the tail (rufous in C. basalis and C. minutillus) and has a pale face without a dark eye streak (dark face with a darker eye streak in C. basalis) and a white throat (gray in C. basalis). Song has rising whistles (falling whistles in C. basalis).

Voice Series of rising whistled notes like those to call a dog, “feee, feee, feee . . .”, given about 2–3/sec, and a descending “pee-eer” alone or at the end of rising whistle series.These calls are 2.5–3 kHz at their loudest and are given at 14–16 per sec. Songs are similar in New Zealand and in eastern and western Australia. (C. l. lucidus and C. l. plagosus) (NSA, RBP) and on Rennell and Bellona Is C. l. harterti) (Mayr and Diamond 2001).Birds in courtship feeding give a quiet “cheep” (Higgins 1999). Begging calls of young, (a) “peep” in nestlings from hatching

412 Shining Bronze-cuckoo Chrysococcyx lucidus to day 5, a short call 0.05–0.07, rising then falling in pitch; (b) “whine” with a single or double band of sound at 8 kHz, maintained on a pitch and lasting 0.3– 0.4 sec, and (c) “ze-ba” buzz of 0.1–0.2 sec (“ze”) followed by short rising “ba” calls of 0.01–0.02 sec, sometimes given in a feed as a series of “ba” notes. Begging calls “peep” and “whine” in Western Australia are similar to those of young of common host Yellow-rumped Thornbill Acanthiza chrysorrhoa, whereas begging calls of cuckoos reared by other hosts are like calls of cuckoos reared by the thornbill (Payne and Payne 1998a).

Range and status Australia and New Zealand birds are long-distance migrants, arriving in September–October in breeding season (Pizzey 1980, Heather and Robertson 1997). Observed in passage across Torres Strait (Fell 1947), regular in passage on Lord Howe I, and a summer resident on Norfolk I (Galbraith and Galbraith 1962, Schodde et al. 1983, Higgins 1999). In Vanuatu, breeding birds are limited to islands with their host species Gerygone flavolateralis (Dutson 2001). In New Zealand these cuckoos are breeding visitors in summer. In Queensland they occur in all seasons, while migrants move north along the coast in autumn and south in spring, including the subspecies plagosus and lucidus (Marchant 1972, Gill 1983b, Blakers et al. 1984, Higgins 1999). A few birds remain in SE Australia, Tasmania, and North I of New Zealand in winter. Few records from NW Australia and none from Northern Territories (Storr 1980, Higgins 1999). Vagrant to Kermadec Islands, The Snares and Auckland Islands (Higgins 1999). In the winter from February and early April to late October, they are in the Lesser Sundas (Lombok, Sumbawa, Komodo, Flores, Sumba, Roti, Dao, Timor, Wetar), New Guinea and nearby islands ( plagosus), and the Bismarck Archipelago, New Caledonia, Bougainville and the Solomon Islands (lucidus) (Rensch 1931, Galbraith and Galbraith 1962, Diamond and LeCroy 1979, Hadden 1981, Mees 1982, Coates 1985, 1991, White and Bruce 1986, Noske 1994,Verhoeye and Holmes 1998, Higgins 1999, Kratter et al. 2001, Mayr and Diamond 2001, AMNH, BMNH) and north to Micronesia

(Kapingamarangi Atoll) (Buden 1998). Common in New Zealand and Australia in wooded country. Breeding densities in Australia, c. 1 bird/km2 (Blakers et al. 1984, Higgins 1999). Numbers have declined in the past 50 years in western Australia, in former woodlands that now are cultivated (Saunders and Ingram 1995). Island birds C. l. layardi on New Caledonia,Vanuatu, Banks I, Santa Cruz Islands, Loyalty Islands, and Rennell and C. l. harterti on Bellona Islands in the Solomon Islands are resident.

Habitat and general habits Forest, woodland, scrub, mainly areas with more than 380 mm of rain; sea level to 1000 m. In winter, it occurs in secondary growth, forest edges, scrub, savanna, village gardens, Casuarina groves, and in forest canopy, occasionally mangroves and pine plantations; from sea level to 700 m on Sumbawa, 500 m on Sumba, 1900 m on New Guinea, 1000 m on Bougainville (Coates 1985, 1991, Coates and Bishop 1997), and 1200 m in New Zealand (Heather and Robertson 1997).They forage in tree and shrub canopy and on the ground, and they join mixed-species flocks of resident insectivorous songbirds (Goodwin 1974, Diamond 1975, Gill 1980a, Higgins 1999).

Food Insects, mainly caterpillars, and beetles, ants, diptera, dragonflies; mollusks, spiders, and eggs of birds (Higgins 1999).

White-eared Bronze-cuckoo Chrysococcyx meyeri 413

Displays and breeding behavior Males are territorial, with one male on a hillside, yet sometimes six or more birds call in a single tree. The adult male feeds the female in courtship, and sometimes he feeds a fledged young cuckoo, perhaps in misdirected courtship feeding (Friedmann 1968, Brooker and Brooker 1989a, Higgins 1999).

Breeding In Queensland from July to early December, in Western Australia from October to December (Storr 1984a,b, 1991, Brooker and Brooker 1989a), in New Zealand October to January (Gill 1982a), in Chatham Islands, where the host also nests before the cuckoo season, the cuckoos breed in November and December (Falla et al. 1978), on Norfolk I in December (Galbraith and Galbraith 1962), and in New Caledonia in September and October (laying females, UMMZ). Brood-parasitic. Hosts, in New Zealand, Little Barrier I and Norfolk I the cuckoos use Grey Gerygone (Grey Warbler) (Gerygone igata) (Gill 1982a, 1983a, Schodde et al. 1983, Cometti 1986); in Vanuatu they use Melanesian Gerygone Gerygone flavolateralis (Dutson 2001); in Chatham Islands the endemic Chatham Gerygone (Chatham Island Warbler) (Gerygone albofrontata) (Falla et al. 1978, Heather and Robertson 1997); in Australia Acanthiza thornbills, Malurus fairy-wrens and Petroica robins rear young cuckoos, which use both covered nests and open nests (Beruldsen 1980, Brooker and Brooker 1989a, 1992, 1994, Payne and Payne 1998). In New Caledonia they use Fan-tailed Gerygone Gerygone flavolateralis (Hannecart and Letocart 1980).

Eggs are .unmarked bronze to olive-green, 18 ⫻ 12 mm, the color and pattern unlike that of any host in Australia; the dark egg is hard to see in a covered nest.The fresh egg is olive; the dark color can be removed with moisture to reveal a pale blue shell (Campbell 1900). Laying females may enlarge the host entrance when they enter the nest, then they usually remove a host egg or two. Laying occurs an hour after sunrise and takes only 18 sec; the cuckoo enters the nest part way, then backs out while removing a host egg in its bill (Brooker et al. 1988). Incubation period 13–14 (to 17) days (the longer periods are in New Zealand, and in nests where the egg was laid before the host began to incubate). Nestling evicts host eggs and young, usually 2 days after it hatches (Courtney and Marchant 1971, Marchant 1986, Brooker and Brooker 1989b, Payne and Payne 1998a), and at 3–7 days in New Zealand where nestling development is slower (Gill 1982b, 1983a). Its eyes open in six days. Nestling period 18–21 days. The young leaves the nest at 22 g, or 88% of adult weight. The young cuckoo is cared for by foster parents for up to five weeks after fledging. Breeding success in Australia, of 38 cuckoo eggs, 13 (34%) hatched and 8 (21%) fledged, the failures due to desertion or because the egg was broken. In Western Australia, of 35 eggs in nests of Yellow-rumped Thornbill, 13 (72%) hatched and 11 (61%) fledged, the failures due to egg burial and egg desertion (Brooker and Brooker 1989b). In New Zealand, of 23 eggs that were observed up to termination, 16 (70%) hatched and 12 (52%) fledged; of the failures 3 were taken by a predator and 1 died from inclement weather (Gill 1983a).

White-eared Bronze-cuckoo Chrysococcyx meyeri Salvadori, 1874 Chrysococcyx meyeri Salvadori, 1874, Annali del Museo Civico di Storia Naturale di Genova, 6, p. 82. (Hatam, Arfak Mountains ⫽ New Guinea] Other common names: Meyer’s Bronze-cuckoo, Mountain Bronze-cuckoo. Monotypic. Other names: Chalcites meyeri (Salvadori).

Description ADULT: Sexes similar, male, crown bright green, back and wing coverts glossy bronze green, wing dark with basal 2/3 of feather rufous on outer edge and the tip gray-brown, tail unbarred green above on central feathers T1,T2 to T5 with white spot on tip, outer tail feathers T5 black with white bars on

414 White-eared Bronze-cuckoo Chrysococcyx meyeri inner vane and marginal notches on outer vane, face green with a white ear patch; underparts white with dark glossy green bars (1-2 mm) from throat to belly, under wing coverts green with narrow (0.8 mm) white bars, underwing P1–P7 with inner vane rufous forming a pale rufous bar in flight; female forehead and forecrown rufous chestnut; eye-ring red (male) to gray (female), iris pale brown to gray-brown, darker on inner rim, bill black, feet blue-gray. JUVENILE: Head gray, back gray-brown, wing graybrown with rufous as in the adult, tail rufous with a subterminal blackish band 8 mm wide and indistinct gray near tip, the tail lacks white spots; underparts unmarked gray; iris brown, bill black, feet light brown. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, MCZ, MSNG, MVZ,YPM.

Measurements and weights Wing, M (n ⫽ 8) 86–91 (88.6 ⫾ 2.0), F (n ⫽ 12) 85–91 (89.1 ⫾ 1.6); tail, M 57–62 (59.8 ⫾ 1.9), F 57–66 (62.7 ⫾ 4.5); bill, M 14.0–15.4 (14.6 ⫾ 0.5), F 14.7–17.0 (15.5 ⫾ 0.6); tarsus, M 13.6–16.4 (14.8 ⫾ 0.9), F 13.6–18.4 (16.3 ⫾ 1.2) (AMNH, FMNH, MCZ,YPM, MSNG). Weight, M (n ⫽ 6) 15–19.5 (17.8); F (n ⫽ 6) 19.0–27.4 (21.2) (AMNH, MVZ, USNM). Wing formula, P8 ⫽ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 15 cm. Small glossy green cuckoo with white ear coverts. Face dark, above deep green (male with green forehead, female with rufous forehead), eye-ring red (gray in female), below white with prominent dark bars, wings rufous.Wing from below has a pale rufous bar in flight (white in Rufous-throated Bronze-cuckoo C. ruficollis), tail is mostly dark below (C. ruficollis has the whitish bars broader than the black bars). The juvenile is gray-brown above (green in C. ruficollis) and gray below (white belly in C. ruficollis).

Voice Song, 5–8 high-pitched clear notes, “peer peer peer . . .” dropping slightly in pitch, the final note following a pause.Also a complex song of four pairs of downslurred notes, the pairs rising and falling (Stein 1936, Beehler et al. 1986).

Range and status New Guinea and Batanta I (Rothschild and Hartert 1907, Hartert 1930, Hartert et al. 1936, Mayr and Rand 1937, Junge 1953, Gyldenstolpe 1955, Gilliard and LeCroy 1961, Parker 1981). Resident. Widespread and fairly common in hills and lower mountains (Coates 2001).

Habitat and general habits Rainforest, monsoon forest, secondary forest, forest clearings and edge, shade trees and gardens. Occur in lower montane habitats from 500 to 1500 m, less common at lower elevations to sea level and rare above 1500 m. Slow-moving in the canopy. They forage in canopy leaves, and they are regular in mixed-species foraging flocks of small insectivorous songbirds especially warblers and flycatchers (Diamond 1972, Coates 1985, 2001, Beehler et al. 1986).

Food Insects, caterpillars in canopy leaves (Coates 1985).

Breeding Females are in laying condition in October (AMNH 301842, Weylands Mts; AMNH 419683, 419864, Mafulu). Presumably brood-parasitic, the hosts and eggs are unknown.

Little Bronze-cuckoo Chrysococcyx minutillus 415

Little Bronze-cuckoo Chrysococcyx minutillus Gould, 1859 Chrysococcyx minutillus Gould, 1859, Proceedings of the Zoological Society of London, 1859, p. 128. (Port Essington, Northern Territory, Australia) Polytypic. Eleven subspecies. Chrysococcyx minutillus minutillus Gould, 1859; Chrysococcyx minutillus poecilurus (G. R. Gray, 1862); Chrysococcyx minutillus misoriensis (Salvadori, 1875); Chrysococcyx minutillus crassirostris (Salvadori, 1878);Chrysococcyx minutillus rufomerus Hartert, 1900; Chrysococcyx minutillus barnardi Mathews, 1912; Chrysococcyx minutillus salvadorii (Hartert and Stresemann, 1925); Chrysococcyx minutillus albifrons ( Junge, 1938); Chrysococcyx minutillus aheneus ( Junge, 1938); Chrysococcyx minutillus jungei (Stresemann, 1938); Chrysococcyx minutillus peninsularis Parker, 1981. Other common names: Malay Bronze-cuckoo, Gould’s Bronze-cuckoo, Rufous Bronze-cuckoo, Dark-backed Bronze-cuckoo, Pied Bronze-cuckoo.

Description ADULT: (C. m. minutillus) male, forehead white, crown bright green, back bronze green, wing bronze brown, outer primaries with brown outer vane and rufous inner vane, tail rufous with white tip and subterminal black bar on inner vane, inner rectrices T1 unmarked rufous bronze to tip,T2 and T3 rufous tip with black subterminal bar,T4 with only white and black subterminal bar, T5 with outer vane bronzebrown and inner vane rufous with black bars and 2 white spots, face pale, white streak over the eye; underparts white with thin green to blackish bars, under wing coverts whitish boldly barred black (dark bars 2 mm wide); conspicuous red eye-ring, iris red, bill black, feet gray. Female, upperparts (especially the crown) less green or bronze glossed, face darker than male; underparts, chin and throat unbarred white, breast barred bronze 1.2 mm dark bars, belly and under tail coverts with wider dark bars (3 mm), less broadly barred than male; eye-ring tan to yellow, greenish or white, iris brown. JUVENILE: Forehead dark without white frosting, head and back gray brown, paler than adult, wing gray brown, tail as in adult, cheek pale; underparts white to rufous with shadow barring on flanks,

under wing coverts unbarred or nearly so; eye-ring brown to dull red, iris brown. NESTLING: (C. m. barnardi) naked at hatching, skin pink to black, sparse yellowish hair-like down on head and back (McGill and Goddard 1979, Brooker and Brooker 1989a). SOURCES: AMNH, BMNH, BPBM, FMNH, KU, MCZ, MSNG, MZB, RMNH, SMTD, USNM, WAM, WFVZ, ZMA, ZMB, ZMS, ZMUC, ZRC.

Subspecies Parker (1981) recognized several species in this complex, whereas Ford (1981) considered them conspecific.The bill width varies within some populations, especially in Borneo. Chrysococcyx minutillus peninsularis Parker, 1981; crown darker green than the back, face with white frosting spots to forehead and around eye; S Thailand and Malay Peninsula; Chrysococcyx minutillus albifrons ( Junge, 1938); forehead and above the eye with white frosting, underparts whiter and less barred; Java and Sumatra; Chrysococcyx minutillus aheneus ( Junge, 1938); upperparts darker bronze than albifrons and with less white on the head, the forehead and supercilium more white in male than in female, bill broad to narrow; Borneo, southern Philippines (Negros, Mindanao), Sulu Islands (Basilan,Tawitawi, Bongao, Sanga Sanga); Chrysococcyx minutillus jungei (Stresemann, 1938); upperparts brownish green with bronze gloss, forehead with few whitish or buff spots; Sulawesi, Peleng, Madu, Flores; Chrysococcyx minutillus rufomerus Hartert, 1900; upperparts dark green with bright bronze gloss, the crown nearly black, the forehead, face and cheek dark, wing sometimes with minute white spots, the tail dark,T1 blackish,T2 to T5 with reduced rufous area compared with other forms, the bars below nearly black; E Lesser Sundas (Banda Sea islands: Romang, Kisar, Leti, Moa, Sermata, Damar,Wetar);

416 Little Bronze-cuckoo Chrysococcyx minutillus Chrysococcyx minutillus crassirostris Salvadori, 1878; upperparts dark blue-green with no bronze gloss, wing dark blue with a large white patch, tail dark with T1 unbarred blue-black,T2 to T4 blackish with white spot at tip on outer web, T5 black with white bars and tip; underparts unbarred white; female dull brown above, wing brown with or without white patch, tail brown marked as in male, face brown or brownish green, the breast and flanks white with indistinct brown bars, center of belly unbarred white; eye-ring red (male and female), iris brown to red, bill black, feet gray to black; juvenile, head and back rufous without metal gloss, wing rufous, tail rufous, T1 with gray subterminal band, T5 with black and white bars on inner web, face rufous; underparts unbarred white; Kai Islands,Tanimbar Islands,Tayandu Islands; Chrysococcyx minutillus salvadorii (Hartert and Stresemann, 1925); male, like crassirostris, small white patch on wing, underparts white with irregular blackish bars;Tepa, Babar Islands; Chrysococcyx minutillus misoriensis (Salvadori, 1875); upperparts bronze green, no white on wing, underparts barred; Biak; Chrysococcyx minutillus poecilurus (G. R. Gray, 1862) (includes C. russata Gould, 1868, and C. poeciluroides Salvadori, 1878; forehead and crown brown, upperparts bronze to greenish bronze, underparts barred, breast often with rufous wash; New Guinea and nearby islands (Seram, Aru Islands, Misool, Dampier Islands, D’Entrecasteaux Islands), Timor, northern Australia (N Queensland from eastern Cape York along coast to Townsville and Ayr); Chrysococcyx minutillus minutillus Gould, 1859; forehead white, crown bright green, back bronze green; northern Australia in Kimberley District, Melville Island, Northern Territory and NW Queensland (Gulf of Carpenteria east to Normanton); Chrysococcyx minutillus barnardi Mathews, 1912; plumage like minutillus but dull gloss above, larger; eastern Australia (SE Queensland and NE New South Wales). The species was earlier known as Chrysococcyx malayanus from the following description of Raffles (1822): “Cuculus malayanus. This species has some affinity to the C. lucidus. It is about seven inches in length: brown above, with a greenish gloss, particu-

larly on the scapulars. The whole under parts are transversely barred with white and brown undulations. The wings are long, extending to about the middle of the tail; the coverts edged with ferruginous. The tail consists of ten feathers, the upper surface greenish brown, and the lower barred with brown, black and white.The bill is somewhat compressed at the base, and the nostrils are prominent. There is a row of white dots above the eyes. Native of the Malay Peninsula.”The description of barring on the tail and the lack of detail on the throat reveals that the type specimen of Cuculus malayanus Raffles, 1822, in fact was a female C. xanthorhynchus (Horsfield, 1821), as suggested by Oates (1887) and Parker (1981). No region is known where two breeding forms occur together and do not interbreed.The songs of all forms that have been recorded or described appear to be the same. Following earlier comments ( Junge 1938, Deignan and Amos 1950, Smythies 1957) on variation in cuckoos in Borneo, Parker (1981) described a thin-billed form, Chrysococcyx minutillus cleis.Although he found thin-billed cleis to have more white on the forehead than thick-billed C. m. aheneus, the distinction does not hold with additional specimens (KU 43141, 43148–50, BPBM 15024, 15084, cf. Thompson 1966; WFVZ 37587–91, cf. Sheldon et al. 2001). In Parker’s sample, female aheneus had bill width at nostril of 5.1-6.2; no females were in his sample of cleis; however, female MZB 24737 has bill width of 3.9. For males, bill width at nostril as measured by Parker is 4.0–4.7 for cleis and 5.0–6.3 for aheneus, bill width at nostril for Thompson’s birds (KU, BPBM) is 5.3, 5.6, 5.7 and 6.0 for four broadbilled adult males and 4.8 and 5.0 for two narrowerbilled females, and bill width at nostril for WFVZ birds is 4.3, 5.3 and 5.3 for males, and 5.1 and 5.3 for females; the narrow-billed male has somewhat less white on the head than the other males (not more, as in Parker’s sample). Because some birds are intermediate in bill form between thin- and thick-billed birds and because these birds are not distinct in plumage, I regard cleis as a synonym of C. m. aheneus. C. m. rufomerus, sometimes called a distinct species “Gould’s Bronze-cuckoo”, is intermediate in tail color and sometimes has small white spots on

Little Bronze-cuckoo Chrysococcyx minutillus 417 the wing, while the plumage color is like the rufous birds of Sulawesi and Borneo. Ford (1981) suggested that the small white spots on the wing and the dark tail suggest introgression with crassirostris, and the plumage color indicates a close relationship with the other C. minutillus. C. m. crassirostris are sometimes considered a distinct species, Pied Bronze-cuckoo C. crassirostris. They differ from other forms in the species complex in being strongly sexually dimorphic and in lacking bars on the underparts in the adult male plumage, in the deep blue-green plumage without a bronze cast on the upperparts, and in having a conspicuous white wing patch in the male; females are slightly bronze above or less often dull green (AMNH 627129), some have a white wing patch (MSNG 13760, AMNH 627129) and others lack the white wing patch (AMNH 627130 has none, in AMNH 627131 only a single covert has a white edge), and females have indistinct bars on the side; juveniles are rufous above with uniformly rufous wings and tail and lack bars on the underparts. C. m. crassirostris and other bronze-cuckoos in the C. minutillus complex of populations are allopatric, although both may occur in the Moluccas and far western New Guinea. The origin of specimens attributed to crassirostris from the Moluccas and New Guinea (Salvadori (1878b, 1881) was questioned because historical records are lacking to support the information on the specimen labels, which are not original field labels (Parker 1981).The other problem is that the northern birds do not all appear to be crassirostris. These northern birds are green and have white in the wing, much as in the type specimen of crassirostris in Genoa (MSNG 13760) from Tual, Kai Islands, and in adult males in AMNH. The adult male called a “syntype” of crassirostris at the Leiden Museum (RMNH 88185) was said to be taken on Halmahera; other birds from the Moluccas in the Leiden series are female or juvenile or are in molt from juvenile to adult plumage. More importantly, although it is thick billed, RMNH 88185 is not crassirostris: the plumage is green above and barred below (not unbarred as in male crassirostris on the Kai Islands), though not as heavily barred as other (non-crassirostris) subspecies including the holotype of misoriensis, and it differs from other populations of

C. minutillus; perhaps it represents an endemic form on Halmahera.Two other crassirostris in the series at Leiden (RMNH 88182, 88183) taken at Sorong in extreme western New Guinea, dark green above with no bronze gloss, with a large white patch on the wing, and white below with no dark bars; may be crassirostris, bright green above but called females on the label.Another cuckoo (MSNG 13763) taken at Warmon near Sorong, with no white in wing, bronze not green above and partly barred below, a juvenile in molt to barred adult plumage, was the type of Salvadori’s (1878) C. m. poeciluroides, now a synonym of C. m. poecilurus which also occurs nearby on Misool and Seram. The occurrence of two kinds of cuckoos in the same region would be reason to consider them two distinct species. Hartert and Stresemann (1925) noted the overlap with an “!” after the “Sorong” locality for crassirostris; the notation “!” was used when an author considered a name or locality to be in error. Stresemann’s skepticism and the lack of supporting details such as original field labels and notes are the bases of Parker’s role as devil’s advocate against these records (Parker 1981). The same arguments could be made against most birds that were collected in the 1800s! Although it was museum practice in early years to prepare specimens as mounts and include the data in pencil on the mount stands, and many mounts were then re-made into study skins without an original label, the specimens were remounted by Finsch who was at Leiden only after 1897 (Ostende et al. 1997), after Salvadori had worked in the Leiden and Genoa museums and described the species (Salvadori 1878b). Salvadori indicated it when he was uncertain of the origin of specimens in this period (as he did in the same publication for Eudynamys (Microdynamis) parva); he indicated no uncertainty about the localities of crassirostris. The ideas that crassirostris specimens were not taken in the Moluccas, and that if they were there in one season (Parker 1981) and not there in another season are problematic, because no collector has ever seen or taken all kinds of birds that are present in a locality, and there is no other evidence that the birds are migratory. The Ternate bird (RMNH 88184) was called a juvenile (Salvadori 1881), but it is barred on the flanks and breast, has

418 Little Bronze-cuckoo Chrysococcyx minutillus no dark green and is not crassirostris; it is somewhat rufous on the breast and appears to be poecilurus. RMNH 88181 from Ambon is a juvenile with a few bright green feathers above and may be the same kind of bird as taken on Halmahera; RMNH 88182 from Sorong is the same but has some greenish-barred feathers on the flanks, and RMNH 88183 from Sorong is green on the back with a brownish head. Both juveniles and adults and birds in molt from juvenile to adult plumage have been taken on the Kai Islands (AMNH, RMNH). That is, these birds occur in all these stages on these islands, and there is no evidence that birds migrate from there to molt in the Moluccas or western New Guinea. Small glossy cuckoos of New Guinea include the forms misoriensis and poecilurus. Salvadori (1878b) described misoriensis from a specimen taken in Misori (⫽ Biak). The holotype (MSNG 13762), an adult female with the crown gray not dark green, is glossed green above with less bronze than poecilurus (includes poeciluroides Salvadori, 1878), but more bronze than crassirostris, in the type series in the Genoa museum. The underparts are barred blackish bronze; the wing lacks white. A bird from Waigeo (AMNH 300704) is barred below, bronze-green above, intermediate in color between misoriensis and poecilurus. Parker (1981) synonymized poeciluroides with misoriensis but did not examine the Genoa museum material. The type specimen of poeciluroides (MSNG 13763) was taken at Warmon; the Warmon River is near Sorong (Mayr and Meyer de Schauensee 1940b). Other New Guinea birds are bronze not green above, and are poecilurus. Australian birds include poecilurus, minutillus and barnardi. The more rufous birds of northeastern Queensland, sometimes called russatus, are indistinguishable in plumage color and size from birds in New Guinea, as noted by Salvadori (1881), Mees (1964), Ford (1981) and Storr (1984b) and in comparisons of museum specimens in the present study. Although New Guinea birds average smaller, and fewer of them are rufous on the breast than birds in Australia, there is much overlap in size and plumage color. A large New Guinea bird (AMNH 223638, wing 95) was taken in the D’Entrecasteaux Islands

in November, far from the breeding range of cuckoos in Australia at a time when southern birds should have migrated to Australia, so the large birds in New Guinea are not necessarily migrants from the south.The western form in Australia, C. m. minutillus, is distinct yet intergrades with rufous birds which occur in the western Cape York area. The intermediate birds in Australia were described and listed by Parker (1981: 40) even though he considered the cuckoos to comprise two species in Australia. Identification of birds in the Banda Islands is problematic: specimens called migrant C. m. minutillus by Mayr (1939) and Parker (1981) appear to Ford (1982) and to me to be C. m. rufomerus, the form that breeds there. Mayr (1939) considered the birds with no white in the wing to be migrants from Australia. The amount of white speckling on the wing is small (not all birds have the white) and these birds do not differ otherwise (AMNH, Ford 1981). The specimen identified by Mayr (1939) as minutillus from the Kai Islands is a juvenile, and it is not possible to identify juveniles in this set of subspecies.The large southeastern birds barnardi are the only Australian C. minutillus known to migrate to New Guinea; barnardi also move north into northern Queensland (Ford 1981, Higgins 1999). In tropical northern and eastern Australia some birds are intermediate in appearance between minutillus and poecilurus (russatus), which are mainly allopatric. These birds may represent introgression between these cuckoo populations (Goodwin 1974, Ford 1981, 1987a). Another idea is that plumage variants represent color phases, although color phases in cuculine cuckoos usually occur in females, and Ford’s (1981) intermediate birds were mainly males. Nevertheless, reports of differences in egg color (unspotted brown in coastal areas, paler brown with fine spotting inland) and nestling color (blackish skin in coastal areas, flesh-color in inland areas) are consistent with the idea that there are two species, poecilurus in the coastal areas and minutillus inland (Nielsen 1996). As all these cuckoos use the same kind of host (flyeaters Gerygone) where the hosts are known, and the songs are the same, the geographic forms appear to be ecological counterparts across their range and they are considered to be a single biological species.

Little Bronze-cuckoo Chrysococcyx minutillus 419

Measurements and weights C. m. peninsularis, Malay Peninsula: Wing, M (n ⫽ 6) 90.7–97.0 (94.2 ⫾ 2.0), F (n ⫽ 2) 83.3–94.0 (93.7) (Parker 1981); wing, M (n ⫽ 7) 92–99 (94.7 ⫾ 3.2), F (n ⫽ 6) 92–100 (95.3 ⫾ 3.1) (AMNH, BMNH, SMF, USNM, ZMUC, ZRC); C. m. albifrons, Java and Sumatra: Wing, M (n ⫽ 10) 90.0–99.0 (94.1 ⫾ 2.8), F (n ⫽ 4) 91.0–95.6 (92.9 ⫾ 2.0) (Parker 1981); Java: Wing, M (n ⫽ 27) 92–99 (94.9 ⫾ 2.1), F (n ⫽ 3) 93–94 (93.3) (MCZ, RMNH, USNM); C. m. aheneus: Wing, M (n ⫽ 8) 90.8–98.3 (93.8 ⫾ 2.5), F (n ⫽ 10) 87.7–96.3 (92.7 ⫾ 2.9) (excluding “cleis”) (Parker 1981); Borneo: Wing, M (n ⫽ 21) 90–99 (93.8 ⫾ 2.6), F (n ⫽ 14) 92–101 (94.3 ⫾ 2.8) (AMNH, BMNH, BPBM, KU, MCZ, RMNH, USNM, WFVZ); Philippines and Sulu Islands: M (n ⫽ 7) 89–97 (92.1 ⫾ 2.8), F (n ⫽ 8) 89–96 (92.1 ⫾ 2.3) (AMNH, DMNH, SMF, UMMZ); C. m. jungei, Sulawesi: Wing, M (n ⫽ 8) 87.6–92.7 (89.7 ⫾ 1.9), F (n ⫽ 3) 87.1–89.3 (89.4) (Parker 1981); wing, M (n ⫽ 15) 86–93 (89.4 ⫾ 2.4), F (n ⫽ 7) 88–93 (91.3 ⫾ 1.9) (AMNH, MZB, RMNH, SMTD, USNM, ZMB); C. m. rufomerus:Wing, M (n ⫽ 14) 95–99 (96.9 ⫾ 1.2), F (n ⫽ 4) 91–96 (93.5) (AMNH, ZMB); C. m. crassirostris, Kai and Tanimbar Islands:Wing, M (n ⫽ 9) 83–97 (90.0 ⫾ 3.9), F (n ⫽ 3) 88–91 (89.7); tail, M 57–60 (58.6 ⫾ 1.2), F 56–61 (59.3); bill, M 14.1–15.4 (14.8 ⫾ 0.4), F 13.4–14.9 (14.6); tarsus, M 14.2–15.4 (14.8 ⫾ 0.8), F 13.4–14.9 (14.3) (AMNH, RMNH); C. m. salvadorii: Wing, M (n ⫽ 1) 97.0 (AMNH); C. m. misoriensis, Biak:Wing, F (n ⫽ 1) 84 (MSNG); C. m. poecilurus, New Guinea (excluding Waigeo and Biak): Wing, M (n ⫽ 18) 83–96 (91.5 ⫾ 2.5), F (n ⫽ 5) 88–97 (91.8); Australia, M (n ⫽ 11) 92–98 (94.8 ⫾ 2.3), F (n ⫽ 5) 90–99 (94.8 ⫾ 3.5) (AMNH, FMNH, MSNG); C. m. minutillus, Australia: Wing, M (n ⫽ 17) 91–99 (95.1 ⫾ 2.2), F (n ⫽ 8) 91–95 (93.0); tail, 57–66 (62.1 ⫾ 3.4), F 57–61 (58.8 ⫾ 1.6); bill, M 15.6–18.4 (17.1 ⫾ 0.7), F 16.6–18.2 (17.7 ⫾ 0.7); tarsus, M 13.1–17.3 (15.7 ⫾ 0.9), F 14.7–17.5 (15.9) (Higgins 1999); Australia: wing, M (n ⫽ 20) 88.1–98.0 (93.6 ⫾ 2.8), F (n ⫽ 8) 90.0–94.0

(92.4 ⫾ 1.4) (Parker 1981); Melville I: Wing, M (n ⫽ 10) 90–98 (92.9 ⫾ 2.3), F (n ⫽ 5) 90–99 (94.8 ⫾ 3.5) (AMNH); C. m. barnardi: Wing, M (n ⫽ 4) 102–106 (104.3), F (n ⫽ 3) 98–99 (98.7) (Parker 1981); wing, U (n ⫽ 12) 94–105.5 (99.4 ⫾ 3.4) (Ford 1981); wing, M (n ⫽ 4) 92–102 (96.3) (AMNH). Weight, C. m. peninsularis: F (n ⫽ 1) 17.7 (Wells 1999); C. m. aheneus: M (n ⫽ 8) 17.5–22 (18.9), F (including 1 laying, 23.0) (n ⫽ 3) 17.5–23.0 (19.4) (Thompson 1966, KU, MZB,WFVZ); C. m. poecilurus: U (n ⫽ 1), 20 (Gilliard and LeCroy 1966); C. m. minutillus: M (n ⫽ 5) 14.5–17.0 (15.3), F (n ⫽ 3) 15.4–17 (16.5) (Higgins 1999); C. m. poecilurus (includes intermediates with C. m. minutillus): M (n ⫽ 5) 16.9– 20.2 (18.5), F (n ⫽ 4) 17–21 (18.8); C. m. barnardi: M (n ⫽ 3) 17–20 (18.7), F ( juvenile, n ⫽ 1) 18.5 (Higgins 1999). Wing formula (C. m. rufomerus, C. m. aheneus C. m.crassirostris), P8 ⱖ 9 ⱖ 7 ⱖ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⱖ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 15–16 cm. Small glossy green to rufous bronze cuckoo with pale face and underparts distinctly barred, black bill, and (males) red eye-ring. C. m crassirostris, small green- to bluebacked cuckoo with white wing patch, unbarred (male) or lightly barred (female) underparts, black bill, and (males and females) red eye-ring; differs from other forms of C. minutillus in Wallacea by white wing patch, lack of glossy bronze on green back and absence of barring on underparts of male, and lightly-barred underparts in female, and by red eye-ring in female as well as male; the juvenile is more rufous than C. minutillus. Female C. minutillus in southeast Asia differs from southeast Asian C. xanthorhynchus, which also has white spots on the face, barred underparts and black, white and rufous in the tail but is mostly rufous in the tail; C. minutillus has the upperparts more green, coarser barring on chin and throat, and a yellowish eyering. In New Guinea and Australia, the adult differs from C. basalis and C. lucidus by the more distinctly barred underparts and red eye-ring, and from C. lucidus by rufous in tail, and the juvenile differs from C. basalis by absence of dark eye streak and

420 Little Bronze-cuckoo Chrysococcyx minutillus from C. lucidus by tail pattern (rufous base of tail in C. minutillus, as in C. basalis). Flight is swift and direct, as in other bronze cuckoos.

Voice Two distinct vocalizations are given. (1) Whistled song, a plaintive “teu teu teu teu”, 4–6 descending whistled notes given in 1.2 sec at about 3 kHz, the notes dropping in pitch from first to last and the time between notes becoming longer. (2) Trill, “te-te-tete-te”, 16 notes in a sec, the trill held at 3 kHz or descending slightly in pitch and lasting 2-4 sec. The calls are similar in southeast Asia (Scharringa 1999, Robson 2000a), Borneo (Holmes and Burton 1987, van Balen and Prentice 1997), Sulawesi (Watling 1983), New Guinea (Beehler et al. 1986) and all three forms in Australia—C. m. minutillus, C. m. poecilurus and C. m. barnardi (Ford 1981, Buckingham and Jackson 1990, Higgins 1999).The trills in Sulawesi (C. m. jungei ) and Timor (C. m. poecilurus) are like trills in Australia. On Tanimbar, crassirostris has a rapid trill of high notes on one pitch, and a descending series of high “pi” notes (Coates and Bishop 1997, B. King recordings), like calls of Indonesian and Australian Little Bronze-cuckoos (Figure 6.6).

Range and status Widespread from the Malay Peninsula through the Greater Sunda Islands, southern Philippines, Lesser Sundas and elsewhere in Wallacea, New Guinea and adjacent islands and northern Australia (Ford 1981, Parker 1981, White and Bruce 1986, Dickinson et al. 1991, Coates and Bishop 1997, Higgins 1999, Wells 1999, Robson 2002). C. m. minutillus are resident throughout the year in the Kimberley and near Darwin, Northern Territory, and in Queensland, and may move within Australia, but there is no compelling evidence of migration of cuckoos out of Australia (Ford 1981, Parker 1981, Johnstone and Burbidge 1991, Higgins 1999). This form or more likely C. m. barnardi seen in migration across the Torres Strait, and large bronze birds in southern Western Division, Papua New Guinea, might be from Australia (Blakers et al. 1984, Storr 1984b, Coates 1985, Beehler et al. 1986, AMNH). C. m. barnardi and C. m. poecilurus are partial migrants within Australia and occur in winter in northern Queensland

north of their breeding range. Birds as large and rufous as breeding glossy-cuckoos in the Cape York peninsula are known elsewhere, with a male at Derby, Western Australia, in July (AMNH 627083) and a female at Timor in August (AMNH 627051). The other forms are resident as far as known. Not threatened as a species, the conservation status of certain island forms is not well known. Population densities are 0.01 to 0.7 birds/ha in the Northern Territories, Australia (Higgins 1999). C. m. crassirostris occurs in the Kai Islands (Tual, Rumadan), Tanimbar Islands (Yamdena, Larat), Tayandu Islands (Kur, Taam), the Moluccas (Halmahera, Ternate, Ambon; Goram specimen from Timor Laut, taken by von Rosenberg, was destroyed in the Darmstadt Museum during the war, Mees 1965), New Guinea (Warmon River near Sorong); Tepa, Babar Islands (White and Bruce 1986, Coates and Bishop 1997, AMNH, MSNG, RMNH). Islands in the south of the range are small, low and coraline; northern localities are mostly high, though the river and island coastlines offer lowland habitats.

Habitat and general habits Lowland forest and forest edge, mangroves, monsoon forest, hill forest, honeymyrtle Melaleuca swamp forest, bamboo in Malay Peninsula, vine thickets, swamp forest, thickets of wattle Acacia and stringybark Eucalyptus tetrodonta, in plantations of Albizia and cocoa, bamboo thickets and tree-lined streets and gardens of towns in Singapore and Sulawesi; generally in dense lowland vegetation, in woodland, scrub and forest edge often coastal. In Borneo they live from sea level to 250 m (Sheldon

Genus Cacomantis 421 et al. 2001), in Flores to 1150 m and in Timor to 300⫹ m (Coates and Bishop 1997) and in New Guinea mainly near sea level, scarce above 500 m (Coates 1985). In western Sulawesi they occur at sea level and higher elevations up to 1800 m in rain forest wherever flyeaters Gerygone are found (Ben King, pers. comm.). In Australia, they are mainly in lowland coastal areas and riverine rainforest. In NE Queensland, minutillus are said to occur in more open habitats, such as woodlands and fringes of riverine forest, where White-throated Flyeater Gerygone olivacea is the main candidate host, and poecilurus (⫽ “russatus”) are in denser habitats such as mangroves, vine thickets and riverine rainforest, where Large-billed Flyeater G. magnirostris is the likely host (Nielsen 1996, Higgins 1999). Sing from high, exposed perch, forage in tree canopy, and take caterpillars in low forest-edge shrubs. Search foliage, feed by perch and pounce sallying, take insects and their larvae from foliage. Flight swift and direct, slightly rising and falling.

Food Insects, mainly caterpillars (Hesperiidae), and beetles (Coccinellidae), ants, bugs (hemiptera), sawfly larvae and flying ants (Sody 1989, Higgins 1999, BPBM).

Displays and breeding behavior Males hold breeding territories where they call, and they chase each other and courtship feed the females (Crawford 1972, Mason 1997, Higgins 1999, Wells 1999).

Breeding In Malay Peninsula the calls are heard all year, eggs appear in March and August and fledglings are tended by their foster parents from April to October (Wells 1999), in Java the cuckoos breed in March (Hartert and Stresemann 1925, Hoogerwerf 1949), in northern Borneo a female C. m. aheneus KU

43141 had an enlarged oviduct in November and laying females were taken in February and April (Sheldon et al. 2001,WFVZ 37587, 37591), in Flores they lay April to June and in September (Verheijen 1964), in eastern Queensland the cuckoos breed in October and November, in northern Queensland from September to February, in Northern Territory from November to March (Storr 1977, 1980, Brooker and Brooker 1989a). A reported “brood patch” (Thompson 1966, Sheldon et al. 2001) refers to the ventral apterium; no brood-parasitic birds have a brood patch. Brood-parasitic. Hosts are Gerygone flyeaters/ gerygones (Sulphur-bellied Flyeater G. sulphurea in Malay Peninsula, Java and Flores, Large-billed Flyeater G. magnirostris in New Guinea and Australia,Whitethroated and Fairy Flyeaters G. olivacea and G. palpebrosa in Australia, other Gerygone species suspected in Wallacea) (MacGillivray 1914, Hartert and Stresemann 1925, Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967, Rand and Gilliard 1967, Ottow and Verheijen 1969, Goodwin 1974, Storr 1980, Brooker and Brooker 1989a, Higgins 1999, Wells 1999). They may parasitize sunbirds as well (Black Sunbird Nectarinia aspasia in New Guinea, Olivebacked Sunbird Nectarinia jugularis in Queensland) (Seaton 1962, Mees 1982). Eggs are olive bronze, 21 ⫻ 14 mm in Java (Hellebrekers and Hoogerwerf 1967); olive brown, 21.9 ⫻ 14.6 mm in Flores (Ottow and Verheijen 1969); olive brown, 20.5 ⫻ 14.7 mm in New Guinea (Mees 1982); and uniform brown (form poecilurus) or bronze with tiny darker spots (form minutillus), 19 ⫻ 13 mm in Australia (Beruldsen 1980). Eggs differ in appearance from the smaller and paler pinkish-white, brown-speckled eggs of the host flyeaters, which use closed, purse-shaped nests. Incubation and nestling periods are unknown. The nestling cuckoo ejects the host eggs. The fledged cuckoo is tended by its foster parents for several weeks after it leaves the nest (Higgins 1999,Wells 1999).

Genus Cacomantis S. Müller, 1843 Cacomantis, S. Müller, 1843, Verhandelingen over de Natuurlijke Geschiedenis der Nederlandsche overzeesche bezittingen, Land-en Volkenkunde, 6, p. 177. Bars on tail

transverse in C. (“Penthoceryx”) sonneratii, bars oblique in other species. Small to medium-sized cuckoos with round nostrils, long tails and usually browns and

422 Pallid Cuckoo Cacomantis pallidus grays in the plumage. Type, by subsequent designation, Cuculus flavus Gmelin ⫽ Cuculus merulinus Scolopi (Salvadori 1881).The genus name Cacomantis (Gr. kakos, evil, ill-boding; mantis, prophet) refers to the “rain bird” in folklore; the birds are supposed to predict ill fortune and bad weather. Cacomantis includes the large Australian Pallid Cuckoo C. pallidus, sometimes regarded as a Cuculus or recognized in a genus by itself, Heteroscenes Cabanis and Heine, 1863; and New Guinea White-crowned Black Cuckoo C. leucolophus, often recognized in a monotypic genus Caliechthrus Cabanis and Heine,

1863. C. leucolophus has been considered a koel related to Eudynamys (as in Mason 1997), but its molecular genetics indicates it is related to the brush cuckoos Cacomantis and particularly to C. pallidus. Both differ from koels in the round, not slit-like nostrils as in Eudynamys. Recognizing all birds as members of a single genus follows the confidence in monophyly of the clade; the subsets of this clade would otherwise indicate recognition of five separate genera (one for flabelliformis and castaneiventris, one for sonneratii, one for pallidus, one for leucolophus, and one for merulinus, passerinus and variolosus). Eight species.

Pallid Cuckoo Cacomantis pallidus (Latham, 1801) Columba pallida (Latham, 1801), Index Ornithologicus, Suppl. lx. (New Holland ⫽ New South Wales) Monotypic. Other names: Cuculus pallidus (Latham, 1801).

Description ADULT: Sexes differ. Male, the head, back and rump uniformly light gray to dark gray, wing gray, nape with white spot (often concealed), tail blackish with white notches along rectrices, T5 with a complete white bar seen from below, other rectrices incompletely barred or with a pale spot along the shaft on underside of feather; face with pale superciliary region, darker gray line through eye extends across ear coverts to side of neck; underparts pale gray, not barred except under tail coverts weakly barred, bend of wing white, under wing coverts gray indistinctly mottled darker gray, under wing with incomplete white patch across base of the inner primaries and outer secondaries; eye-ring yellow, iris dark brown (some with a whitish or yellow outer ring), bill black, feet gray. Female, the head dark gray variably streaked light rufous, nape with white spot, mantle light rufous with large black splotches, back brownish gray with variable rufous bars, wing gray brown with buff and white spots on upper coverts and flight feathers dark gray with buff notches, tail as in male, face gray with a whitish forked streak extending from in front of eye to over the eye and auricular and a lower streak below the eye, the throat and breast light gray with black and rufous bars and scallops, the barred region varying from none to a broad band,

belly and under tail coverts lighter gray indistinctly barred medium gray; bill black to greenish black above and with a yellowish green base below. The variation among females in plumage is continuous rather than dimorphic with dark- and light-phases and it appears to be independent of age. JUVENILE: Whitish bird with black face and breast: crown, back and rump white with black streaks, scapulars black with whitish tip, the region forming a dark streak, upper wing coverts white with black streaks. wing gray, the flight feathers with buffy notches and a white edge along the tip, tail blackish with white notches; the face darker than top of head and nape, the forehead and lores dark brown, a conspicuous black streak from eye to bend of wing outlined by white of the crown and superciliary region and by a white streak below the eye; the throat and upper breast blackish, the feathers indistinctly edged white, the lower breast and belly white streaked with dark gray; eye-ring narrow and pale, iris gray and with age becoming ringed with brown, bill gray with pale cutting edges, gape yellow, feet pale gray. The black and white juvenile lacks the rufous of the adult female. NESTLING: Naked, blind at hatching, skin color not described, back has a concave surface (Campbell 1900, Mathews 1918). SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, ROM, UMMZ, USNM,WAM.

Pallid Cuckoo Cacomantis pallidus 423

Measurements and weights Wing, M (n ⫽ 69) 179–204 (193), F (n ⫽ 48) 182– 204 (191.5); tail, M (n ⫽ 69) 151–177 (163.4), F (n ⫽ 45) 151–178 (160.6); bill, M (n ⫽ 61) 23.8– 29.3 (26.9), F (n ⫽ 47) 24.8–29.1 (26.8); tarsus, M (n ⫽ 24) 19.5–25.3 (21.4), F (21) 20.2–25.7 (22.2) (Higgins 1999);Western Australia, wing, M (n ⫽ 10) 181–199 (188.2 ±6.7), F (n ⫽ 5) 181–198 (190.0 ⫾ 7.6) (WAM). Weight, M (n ⫽ 22) 64–118 (89.5), F (n ⫽ 18) 63–106 (85.6) (Higgins 1999). Wing formula, P8ⱖ 9⬎ 7⬎ 6⬎ 5⬎ 4⬎ 10⬎ 3⬎ 2⬎ 1.

Field characters Overall length 3l–32 cm. Large cuckoo with unmarked underparts, gray head with blackish streak in front and behind the eye, and male has a song with rising scale of whistles. Female has crown blackish streaked rufous and the mantle and breast are often marked rufous and black. Juvenile is whitish with black throat and breast. In flight, it often glides like a falcon (small rounded head, long pointed wings and long rounded tail), and it perches in the open and is mobbed by small birds.

Voice Song, a whistled series of c. 8 notes ascending the scale, second note often lower. Other calls, a coarse crescendo call “gahgahGAHGAH” in defense of territory, a three-note call “toy-it-yer, toy-it-yer” in flying towards a female, and a soft “peep” in courtship feeding. Female, a hoarse whistle “kheer”. Begging call of young, a thin piping “chirrup, chirrup” that varies with hunger or with host species (Condon 1947, Slater 1971, Pizzey 1980, Slater et al. 1986, Buckingham and Jackson 1990, Higgins 1999, RBP recordings).

History Life history was summarized in 1865 by Gould, who noted that it was a migrant and a brood parasite. Pallid Cuckoo has often been considered a Cuculus. Australian and other ornithologists have recognized it as Cacomantis pallidus (e.g., North 1899, Bonaparte and Gould, in Mathews 1918), while Mathews considered it intermediate between Cuculus and Cacomantis and described a separate genus Heteroscenes (Cabanis

and Heine, 1863); the name indicating the variation in plumage in the different scenes of life. Rising song is similar to rising song of Cacomantis species, and its three-note call may be the cadence call of other Cacomantis. The plumage lacks the barred underparts of Cuculus, and the juvenile ventral plumage is streaked as in some other Cacomantis. Molecular genetic analysis indicates that C. pallidus is more closely related to other Cacomantis than to Cuculus.

Range and status Australia and Tasmania. Resident, migrant or with erratic movements in most of Australia where birds occur through the austral winter. In northern Australia they occur in all months; in drier parts of Australia mainly a wet-season visitor (Goodwin 1984, Storr 1980, 1984b, Johnstone and Storr 1998). In SW Western Australia adults arrive in May and June and depart in October and November, and juveniles may remain with their foster parents until February (Storr 1991). Migratory movements are seen along coastal eastern Australia, in the MurrayDarling region, at Murphy’s Creek, and on Ashmore Reef and islands off NW Australia (Blakers et al. 1984, Storr 1984b, Higgins 1999). Rare during austral winter in Moluccas (Ternate), Flores and Timor (Hellmayr 1916, White and Bruce 1986) and New Guinea, one record in Torres Straits (Mees 1982, Coates 1985). In Tasmania, they occur only in the breeding season.Vagrant in New Zealand, and accidental on Lord Howe I, Norfolk I, Christmas I and Macquarie I. The distribution of wintering birds through Australia north of 20°S, the reporting rates in summer and winter north of 26°S, and the few sightings elsewhere suggest that most birds remain

424 White-crowned Cuckoo Cacomantis leucolophus in Australia (Storr 1977, 1984a, Blakers et al. 1984, Higgins 1999). Common, the cuckoos occur in residential and suburban areas where honeyeater hosts are attracted to flowering gardens, and in scrub and open woodland. Population density, 1 territory/ km2, locally higher (Loyn 1980, Higgins 1999).

Habitat and general habits Open forests, woodlands, scrub, Eucalyptus mallee shrubland with sparse understory, spinifex plains, mangroves, gardens, more numerous in drier country than in forest. The cuckoos occur in the arid interior parts of Australia and along the coast, both habitats that have high densities of their hosts, the honeyeaters. They feed by peering through shrub and trees and sally-pouncing from an elevated perch, and feed on the ground.

Food Insects, mainly grasshoppers, beetles, caterpillars and other larvae, cicadas; centipedes, millipedes; occasionally tadpoles, seeds and berries ( Johnstone and Storr 1998, Higgins 1999).

Displays and breeding behavior Male courtship feeds the female cuckoo (Kikkawa 1968). In copulation, the female called and a male flew in and mounted, in another pair the male and female rotated tails for 1-2 min (Klapste 1981, Higgins 1999). Adult cuckoo (male?) sometimes feeds a fledged young cuckoo, at times when the young was attended by its foster parents (Cooper 1958, Kikkawa and Dwyer 1962, Watts 1993, Nicholls 1998), the male perhaps responding to the young cuckoo as if it were a breeding female. Females breed in the first year, with a bird in partial juvenile plumage having an egg in the oviduct (Higgins 1999).

Breeding North of 20°S it breeds from March to June and in September, in SW Western Australia from August to October, in eastern Australia from September to December (Higgins 1999). Brood-parasitic. Hosts are mainly honeyeaters with open nests, and shrike-thrushes and cuckooshrikes also with open nests. A total of 111 host species have been reported, and 21 of the 32 hosts that are known to rear the cuckoos are honeyeaters (Brooker and Brooker 1989a, Storr 1991). Cuckoos also lay in the nest of birds that are not successful in rearing them, with hatching but later starvation of nestling cuckoos in the nests of finches (Red-browed Finch Neochmia temporalis) that feed seeds to their young (Cole 1908, Mathews 1918). Eggs are pink, unspotted or with brown specks around large end, 25 ⫻ 16 mm (North 1899, Mathews 1918, Beruldsen 1980).The female cuckoo often removes an egg of the host. Incubation period, 12–14 days. Nestling cuckoo evicts eggs and nestlings of host, usually 1–2 days after it hatches, but in deep nests with large host nestlings it evicts as late as 5 days after hatching. Nestlings sometimes cover and smother the host young.The nestling period is unknown. The young cuckoo is fed by the foster parents for as long as six weeks after it leaves the nest. It is sometimes fed by neighboring birds as well as by the birds that reared it, and occasionally a fledged young visits a nest with young and takes the food brought to the nestlings, or even takes over a nest by throwing out the nestlings (Mathews 1918, Leach 1929, Kikkawa and Dwyer 1962, Hardy and Hardy 1973, Brooker and Brooker 1989a, Smith 1989, Hobbs 1990, Higgins 1999).

White-crowned Cuckoo Cacomantis leucolophus (S. Müller, 1840) Cuculus leucolophus S. Müller, 1840, Verhandlingen over de Natuurlijke Geschiedenis der Nederlandsche overzeesche zebittingen, Land-en Volkenkunde, 1, p. 22 (Lobo Bay, New Guinea) Other names: Caliechthrus leucolophus (S. Müller, 1840).

Monotypic. Other common names: White-crowned Koel.

Description ADULT: Sexes alike, plumage glossy black with a broad white crown stripe, wing black, tail black, the

White-crowned Cuckoo Cacomantis leucolophus 425 rectrices narrowly tipped white, bend of wing with white bars, under wing coverts black with white bars; iris dark brown, bill black, feet dark gray. JUVENILE: Plumage black, without gloss and softer in texture than adult, white crown stripe, wing slate gray, tail slate gray with trace of white tip, rectrices narrower than in adult; underparts, throat slate black barred with white, breast to belly black; iris dark brown. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, MCZ, MSNG, MVZ, RMNH, ROM, SMTD, ZMB.

No anatomical materials (skeleton, spirit specimen) are available for this cuckoo.

Measurements and weights

Range and status

Wing, M (n ⫽ 8) 167–182 (172.3 ⫾ 5.2), F (n ⫽ 6) 165–178 (168.5 ⫾ 4.8); tail, M 162–180 (169.6 ⫾ 7.0), F 159–170 (163.8 ⫾ 3.6); bill, M 24–29 (26.5 ⫾ 2.9), F 22.6–27.8 (25.5 ⫾ 2.2); tarsus, M 23.2–25.0 (24.0 ⫾ 0.7), F 20.6–24.5 (22.5 ⫾ 1.3) (AMNH). Weight, M (n ⫽ 3) 106.5–125 (114.8), F (n ⫽ 1) 99.0 (AMNH, ZMB, Diamond 1972). Wing formula, P7 ⬎ 8 ⬎ 6 ⬎ 5 ⫽ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

New Guinea and Salawati I. Resident.A relationship to Koel Eudynamys is suggested by its black plumage and by the occurrence of white on the head in male koels E. scolopacea of the form “facialis” in Sula Islands.The songs of White-crowned Black Cuckoo are more like songs of the brush cuckoos Cacomantis. Common in hill forest; noted on the Arfak Peninsula (⫽ Vogelkop) and most easily located in Varirata NP, an hour’s drive from Port Moresby.

Field characters

Habitat and general habits

Overall length 33 cm. Large black cuckoo with a white crown stripe. Slow moving and hard to see, in the field it is recognized by its calls.

Forest and second growth at forest edge; mainly in hill country from sea level to 1000 m, uncommonly at higher elevations to 1500–1740 m; their lowland sites are often near hills.The cuckoos live in treetops and the middle story and feed in the canopy. Slowmoving, often seen with a characteristic dipping flight as they cross a clearing. Occur singly or in pairs, occasionally in small groups (Rand and Gilliard 1967, Diamond 1972, Coates 1985, Gregory 1995).

Voice Loud, mournful repetitive song, three clear whistles in a descending series, “too too too”, each series higher in pitch and more insistent, ending with an explosive “week!”, a short, rapid 3–4 note call, suggesting laughter:“ka-ha-ha(-ha)”, and a single burry call “whurr” (Ogilvie-Grant 1915, Mayr and Rand 1937, Coates 1985, 1991, Beehler et al. 1986, NSA).

History Often regarded as the only species in a genus Caliechthrus, it is most similar in molecular genetics to Pallid Cuckoo Cacomantis pallidus. The songs of C. leucolophus with clear whistles in a descending series and with the rise in pitch from phrase to phrase resembles that of other Cacomantis cuckoos.

Food Mainly large insects, including caterpillars, other arthropods, and fruit (D’Albertis and Salvadori 1879, Coates 1985).

Breeding Presumably brood-parasitic.A female was taken with an enlarged ovary in October (SMTD C-29.916). The eggs and other details of breeding biology are unknown.

426 Chestnut-breasted Cuckoo Cacomantis castaneiventris

Chestnut-breasted Cuckoo Cacomantis castaneiventris (Gould, 1867) Cuculus castaneiventris Gould, 1867, The Annals and Magazine of Natural History, (3), 20, p. 269. (Cape York district, Queensland, Australia) Polytypic. Three subspecies. Cacomantis castaneiventris castaneiventris (Gould, 1867); Cacomantis castaneiventris arfakianus Salvadori, 1889; Cacomantis castaneiventris weiskei Reichenow, 1900.

Description ADULT: Sexes alike, head and back slate gray, wing coverts slate gray, wing slate gray, tail slate gray, rectrices notched white; chin dark gray, underparts bright chestnut, underside of tail, inner web of T1 to T5 with narrow white bars, bend of wing white, under wing coverts rufous, same color as breast and belly, under wing slate with a broad white band across primaries and secondaries; eye-ring yellow, iris gray to brown, bill black with yellow gape, feet yellow. JUVENILE: Head and back unbarred dark graybrown, rump and upper tail coverts more rufous (less gray) than in adult, tail dark gray brown, T1 to T5 with rufous edge (narrow rufous edge in T5),T3 to T5 barred black and rufous, tail below rufous gray, face gray-rufous; underparts with belly unbarred brown (C. c. castaneiventris), gray-brown faintly barred darker gray (C. c. weiskei) or whitish buff (C. c. arfakianus), under tail coverts buff to rufous gray, wing band seen from below as in adult; eye-ring narrow, brownish gray, iris dark brown, bill black with yellow gape, feet dull pink. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, FMNH, MCZ, MSNG, ZMB, ZMUC.

Subspecies Cacomantis castaneiventris castaneiventris (Gould 1867); as above, underparts unbarred brown to buff, juvenile as above; Australia (east coast of Cape York and NE Queensland) and Aru Islands; Cacomantis castaneiventris arfakianus Salvadori, 1889; underparts chestnut, smaller; western Papuan

islands, New Guinea from Vogelkop peninsula of Irian Jaya to Weyland and Snow Mountains; Cacomantis castaneiventris weiskei Reichenow, 1900; underparts chestnut; southern New Guinea east to Fly River. No consistent color differences occur between adult arfakianus and weiskei, although juvenile arfakianus are more whitish below than juveniles of the other two forms (Hartert 1925, Gilliard and LeCroy 1961; AMNH, ANSP, FMNH, MSNG).

Measurements and weights C. c. castaneiventris; Australia:Wing, M (n ⫽ 9) 104– 114 (110.2 ⫾ 3.3), F (n ⫽ 2) 111–114 (112.5); tail, M 104–124 (115.2 ⫾ 6.6), F 114, 114; bill, M 20.7–23.4 (21.6 ⫾ 0.9), F 19.9–21.7 (20.8); tarsus, M 17.1–19.8 (18.0 ⫾ 0.8), F 15.1–18.2 (16.7) (Higgins 1999); C. c. weiskei: Wing, M (n ⫽ 15) 102–115 (112.1 ⫾ 3.9), F (n ⫽ 11) 102–118 (112.3 ⫾ 5.0); tail, M 108–131 (116.9 ⫾ 5.9), F 104–128 (115.6 ⫾ 7.7) (AMNH, ANSP, FMNH, MCZ); C. c. arfakianus: Wing, M (n ⫽ 11) 108–117 (112.5 ⫾ 3.3), F (n ⫽ 9) 102–116 (109.6 ⫾ 6.0); tail, M 102-12712.9 ⫾ 6.7), F 102–122 (111.0 ⫾ 6.8) (AMNH, FMNH, MSNG). Weight, C. c. castaneiventris, Australia: M (n ⫽ 4) 26–34 (30.0), F (n ⫽ 1) 32 (Higgins 1999); C. c. weiskei: M (n ⫽ 4) (35.8), F (n ⫽ 3) 35–38 (36.7) (Diamond 1972, AMNH); C. c. arfakianus: M (n ⫽ 6) 27–36 (32.0), F (n ⫽ 5) 31–45 (35.2) (AMNH, ANSP, MCZ). Wing formula, P8 ⬎ 7 ⱖ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 10 ⱖ 1.

Field characters Overall length 24 cm. Small cuckoo with upperparts dark gray and underparts bright chestnut. Color of upperparts similar to that of larger Fantailed Cuckoo C. flabelliformis except upper tail is dark (boldly notched white in C. flabelliformis); underparts with slate chin and the breast and belly more rufous than in Brush Cuckoo C. variolosus infaustus and C. flabelliformis excitus. Juvenile is

Fan-tailed Cuckoo Cacomantis flabelliformis 427 rufous above, paler buff below, unbarred (barred in juvenile C. variolosus and C. flabelliformis).

Voice Song, a loud three-note mournful whistled “seeito-saai”, the first note slurred up, the second shorter, the third slurred up and trilled, c. 2 kHz and lasting 2-3 sec, the cadence slower than the rising “where’s the tea?” song of Brush Cuckoo C. variolosus. A second song is a trill given at 14 notes/sec and descending slightly pitch between 3 and 2 kHz. Call, a single whistled “chir-rip” (Mayr and Rand 1937, Slater 1971, Diamond 1972, Pizzey 1980, Coates 1985, 2001, Beruldsen 1990, Buckingham and Jackson 1990, Gregory 1995, Higgins 1999).

Range and status New Guinea and the western Papuan islands,Yapen I, Aru Islands, Australia (Cape York peninsula, NE Queensland). Resident, observed around the year in Cape York (Rothschild and Hartert 1907, Hartert 1925a, Hartert et al. 1936, Blakers et al. 1984, Storr 1984b, Coates 1985, 2001, Strahan 1994, Higgins 1999).

Habitat and general habits Interior montane forest, tropical rainforest and scrub along river banks, monsoon forest, usually in tree canopy and middle levels. Rare at sea level, mainly hills to 1800 m, locally in lowlands to 2500 m (Coates 2001), in eastern New Guinea often below the altitude of Fan-tailed Cuckoo C. flabelliformis (Diamond 1972). They move through tree canopy feeding on insects, watch from a low perch, fly to the ground, and hover low as they pick insects from

leaves in forest (Beehler et al. 1986, Pizzey and Knight 1998).

Food Insects, including caterpillars and beetles, and ticks (Gilliard and LeCroy 1961, Goodwin 1974).

Displays and breeding behavior Territorial, the male advertises by song. Usually seen alone or in twos.

Breeding In New Guinea (Weyland Mts, AMNH) a female was in laying condition in September, in Australia song is heard from June to August or September. Brood-parasitic. Hosts include Grey-green Scrubwren Sericornis arfakianus (feathered nestling cuckoo, FMNH 280048) in New Guinea, and Tropical Scrub-wren S. beccarii in Australia. Eggs are white freckled reddish, 20.8 ⫻ 14.7 mm (Campbell 1900, Beruldsen 1980, Brooker and Brooker 1989a, Higgins 1999). Incubation and nestling periods are unknown.

Fan-tailed Cuckoo Cacomantis flabelliformis (Latham, 1801) Cuculus flabelliformis Latham, 1801, Index Ornithologicus Suppl. xxx. [Sydney area, New South Wales]. Polytypic. Six subspecies. Cacomantis flabelliformis flabelliformis (Latham, 1801); Cacomantis flabelliformis pyrrophanus (Vieillot, 1817); Cacomantis flabelliformis

simus (Peale, 1848); Cacomantis flabelliformis schistaceigularis Sharpe, 1900; Cacomantis flabelliformis meeki Rothschild and Hartert 1902; Cacomantis flabelliformis excitus Rothschild and Hartert, 1907. Other common names: Fan-tailed Brush Cuckoo.

428 Fan-tailed Cuckoo Cacomantis flabelliformis

Description ADULT: Sexes similar, head and back slate gray, wing slate gray, tail gray with a square tip, the feathers barred white or notched white on inner and outer webs, white notches on outer edge of rectrices, T5 with white bars across inner web and white notch on outer web; underparts, throat gray, breast buff to light rufous, belly and under tail coverts whitish, unmarked or with fine indistinct gray bars, the tail seen from below with white tips and whitish bars, bend of wing white, under wing slate with a broad white band across primaries and secondaries, under wing coverts dark buff. Males are darker and more rufous on breast than females, rufous to belly and under tail coverts and little vermiculate barring on flanks; females have the rufous underparts paler and indistinctly barred on belly and flanks. Eye-ring yellow, iris brown, mouth lining yellow-to red-orange, bill black, feet olive-yellow. JUVENILE: Head, back and wing brown with dull rufous streaks (rusty slate), rump slate with indistinct rusty bars on tips of feathers, tail with prominent white bars or notches, face slate with dark rusty bars; underparts, throat and upper breast mottled rusty and blackish, belly whitish to buff mottled black, paler than the breast, under wing coverts buff sometimes with black bars; mouth lining yellow to orange, iris dark brown, eye-ring greenish yellow, feet brownish yellow.

Cacomantis flabelliformis excitus Rothschild and Hattert, 1907; plumage darker, nearly black above, breast chin and throat gray, breast and belly rufous slate, under tail coverts rufous, tail of adults with little white, tail of juveniles with prominent white or brown bars, juvenile nearly black above; mountains of New Guinea; Cacomantis flabelliformis schistaceigularis Sharpe, 1900; plumage darker chestnut than simus, tail of adults with little white, tail of juveniles with prominent white bars;Vanuatu (New Hebrides); Cacomantis flabelliformis pyrrophanus Vieillot, 1817; underparts rich chestnut, chin and throat rufous or mixed gray and rufous, tail with little white, under wing coverts rufous, unbarred; New Caledonia and Loyalty Islands; Cacomantis flabelliformis meeki Rothschild and Hartert 1902; similar to pyrrhophanus but chin and throat gray; Solomon Islands (Ysabel, Rennell and Bellona Is); Cacomantis flabelliformis sinus Peale, 1848; smaller, chin to breast and belly pale chestnut, rectrices barred; also a melanistic phase “infuscatus” with slate gray plumage, the rectrices unbarred, very little white or brown in tail of juvenile or adult; Fiji. A breeding female taken by C. Filardi in 2004 on Kulombangra I, Solomon Is, is distinct in plumage and is smaller (AMNH, wing 116, tail 127) than other cuckoos of this species.

History

NESTLING: Hatching naked, skin pinkish, becoming dark brown in a few days; edge of gape yellow, palate pinkish orange.

Mason (1982) reviewed the species names and noted that flabelliformis has priority over pyrrhophanus. Stresemann (1924b) determined that infuscatus is a color form of simus.

SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, ROM, SMF, UMMZ, USNM,WAM, YPM, ZFMK, ZMUC, ZSM.

Measurements and weights

Subspecies Cacomantis flabelliformis flabelliformis Latham, 1801; above plumage medium gray, underparts pale gray with pale rufous neck and breast, under tail coverts light rufous, the male unbarred, the female indistinctly barred below; eastern Australia from Cape York to SE Australia, SW Western Australia and Tasmania;

C. f. flabelliformis, Eastern Australia: Wing, M (n ⫽ 42) 138–151 (144.4 ⫾ 3.4), F (n ⫽ 18) 134–147 (140.6 ⫾ 3.8); tail, M 130–149 (140.5 ⫾ 4.3), F 123–151 (135.4 ⫾ 6.4); bill, M 18.2–23.5 (21.5 ⫾ 1.2), F 19.0–22.6 (21.4 ⫾ 1.1); tarsus, M 18.0–22.2 (19.7 ⫾ 1.0), F 18.3–20.9 (19.6 ⫾ 0.6) (Higgins 1999);Western Australia:Wing, M (n ⫽ 16) 135–148 (142.8 ⫾ 3.7), F (n ⫽ 10) 131–143 (136.6 ⫾ 4.4); tail, M 130–148 (141.0 ⫾ 5.9), F 127–143 (136.8 ⫾ 4.7) (WAM);

Fan-tailed Cuckoo Cacomantis flabelliformis 429 C. f: excitus, New Guinea: Wing, M (n ⫽ 8) 138–145 (141.4 ⫾ 2.4), F (n ⫽ 8) 137–143 (140.4 ⫾ 2.1); tail, M 138–151 (145.0 ⫾ 4.7), F 135–151 (142.0 ⫾ 5.9) (AMNH,YPM); C. f. meekei, Solomon Is, wing, M (n ⫽ 2) 140–145, F (n ⫽ 2) 140–142; tail, M 134–148, F I 32–156 (AMNH); C. f. pyrrhophanus, New Caledonia and Loyalty Islands: Wing, M (n ⫽ 12) wing, M (n ⫽ 14) 135–149 (141.6 ⫾ 3.8), F (n ⫽ 4) l39–146(142.5); tail, M 138–158 (148.9 ⫹ 6.l); F, 138158 (147.8) (YPM, ZFMK); C. f. schistaceigdaris, Vanuatu: Wing, M (n ⫽ 6) 128–138 (132.7 ⫾ 3.2), F (n ⫽ 6) 127–137 (130.8 ⫾ 3.9); tail, M 132–138 (134.2 ⫾ 3.0), F 127–137 (131.0 ⫾ 3.5) (AMNH); C. f. simus, Fiji Is: Wing, M (n ⫽ 8) 128–137 (131 ⫾ 2.6), F (n ⫽ 3) 128–130 (128.7); tail, M 135–144 (140.7 ⫾ 3.4), F 130–143 (140) (AMNH). Weights, C. f. flabelliformis: M (n ⫽ 32) 42.6–65 (49.8 ⫾ 4.4), F (n ⫽ 7) 45–53.5 (49.9 ⫾ 3.1) (Higgins 1999), C. f pyrrhophanus, M (n ⫽ 6) 43.5–55.8 (43.7) (AMNH). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 1 0 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 26 cm. Adult is gray above, the tail with white notches on side, below pale rufous with a paler belly and tail with square tip (more rounded in Brush Cuckoo C. variolosus), the tail with white tips and whitish bars (bars incomplete in C. variolosus). Juvenile, brown above with rufous edges to feathers, underparts indistinctly barred gray, tail notched white and rufous above, barred below.

Voice Male gives a descending trill “peeeeer”, about 15 notes in a second, and a rising whistle “p-whee”, each phrase lasting c. 0.8 sec and sometimes given in pairs with a short note between the two phrases, suggesting “cadence call” of other Cacomantis species. Female gives a loud shrill “chireee” (a vibrant and higher-pitched “p-whee”) in courtship feeding and in response to other females (Mayr and Rand 1937, Pizzey 1980, Marchant and Höhn 1980, Buckingham and Jackson 1990, Higgins 1999). Birds in Vanuatu call like birds in Australia (Doughty 1999).

Fledged young cuckoos give a continuous plaintive “zeep-zeep-zeep”.

Range and status Australia, New Guinea (widespread, generally uncommon in the mountains in Arfak Mts,Weyland Mts, Central Ranges and Sattelberg Mts of Huon Peninsula), Solomon Is, New Caledonia and the Loyalty Islands, Vanuatu (New Hebrides), and Fiji (Hartert et al. 1936, Clunie and Morse 1984, Coates 1985, 1991, Beehler 1986, Higgins 1999). Resident in the tropics, elsewhere partially migratory.Tasmanian birds move to Australia. Within Australia, distribution is seasonal in some areas with movements northward and inland for the winter, in Queensland the birds wintering north of the breeding range are mainly immatures. C. f. flabelliformis is a vagrant or passage migrant in interior Australia, occasional in Aru Islands, in New Guinea in southern lowlands east to Vanapa River and E-C mid-montane valleys, vagrant on Lord Howe I and New Zealand. C. f. pyrrhophanus, formerly said to breed in New Caledonia and winter in the Loyalty Islands, occur in each area in all seasons. Other subspecies are resident. In Vanuatu the distribution suggests local fluctuations with island populations going extinct for decades before the islands are recolonized (Diamond and Marshall 1977). Australian birds migrate at night. Common in much of the range; this species is the most common cuckoo in parts of eastern Australia. In Western Australia, it is rare or uncommon where it was common 50 years ago in areas that were woodlands and now are cultivated (Saunders

430 Banded Bay Cuckoo Cacomantis sonneratii and Ingram 1995). Population densities in eastern Australian eucalypt woodland, 0.02-1.0 bird /ha (Blakers et al. 1984, Keast et al. 1985, Higgins 1999).

Habitat and general habits Mainly tall open forests, rainforest, open woodlands, scrub along water courses, acacia thickets, roadsides, mangroves, heath, pastures, orchards; in New Guinea it occurs in the mountains at 1300–3900 m (Beehler et al. 1986, Storr 1991, Pizzey and Knight 1998, Coates 2001). Forage in foliage, more often feed on the ground, feed on tree trunks, and hawk flying insects. Most movements in feeding were sallypouncing, the bird taking food in one attempt in 2 min (Pizzey 1980, Clunie and Morse 1984, Recher and Holmes 1985, Higgins 1999).

Food Insects, mainly hairy caterpillars, and beetles, cockroaches, diptera larvae, flying moths, wasps and ants; spiders and centipedes, also earthworms ( Johnstone and Storr 1998, Higgins 1999).

Displays and breeding behavior Territorial, male advertises by song. Sometimes groups of sexually active birds call within a few meters of each other. Male courtship feeds the female, mounts on her back, steadying himself with wing movements, and copulation lasts about 4 sec, male dismounts and flies off, the female follows him (Smithers 1977, Marchant and Höhn 1980, Smedley 1983). Fledgling is sometimes fed by adult cuckoo (Ambrose 1987, Higgins 1999), perhaps in misdirected courtship feeding.

Breeding In Australia they breed from August to December (Beruldsen 1980, Marchant and Höhn 1980, Pizzey 1980, Brooker and Brooker 1989a, Pizzey and Knight

1998), in highland New Guinea they breed from late April (a fledgling was found on 5 June) to June (Mayr and Gilliard 1954, Diamond 1972), in the Loyalty Is birds call and sing much of the year but there are no breeding records (AMNH), in the Solomon Is on Kulombanga a laying female was taken in May (C. Filardi) and in Fiji they breed from November to February (Clunie and Morse 1984). Brood-parasitic. Hosts, mainly birds with covered or domed nests, especially nests built on or near the ground, include Large-billed Scrub-wren Sericornis nouhuysi in New Guinea (Diamond 1972, Coates 1985); Scarlet Robin Petroica multicolor in Vanuatu (New Hebrides), and covered nests of fairy-wrens Malurus, thornbills Acanthiza and scrubwrens Sericornis and less often the open nests of honeyeaters and flycatchers in Australia (where 17 host species are known) (Storr 1984b 1991, Brooker and Brooker 1989a, Jansen 1990, Strahan 1994, Pizzey and Knight 1998), a honeyeater Myzomela, a whistler Pachyceplala, a triller Lalage, and/or a monarch Myiagra in New Caledonia (Hannecart and Letocart 1980), and Fiji Warbler Vitia ruficapilla in Fiji (Clunie 1973). Australian White-browed Scrub-wrens Sericornis frontalis often build under the fallen bark at the base of Eucalyptus gum trees (Whittingham et al. 1997,Whittingham and Dunn 1998), and the cuckoos use these well-covered subterranean nesting hosts. Eggs are white freckled brown and purple at large end (Australia, Fiji), 21 ⫻ 15 mm (Beruldsen 1980;WAM). Incubation period is less than 13 days. The nestling evicts the host eggs and young on day 2, its eyes open on day 5, and it is fully feathered by day 15. It fledges in 16–17 days, and it may enlarge the nest hole with its body when it leaves the nest. The young cuckoo is tended for 3–4 weeks after it fledges (Barrett 1905, Brooker and Brooker 1989a, Higgins 1999). Breeding success, for 23 cuckoo eggs of known outcome, 10 (43.5%) hatched and 5 (22%) fledged (Higgins 1999).

Banded Bay Cuckoo Cacomantis sonneratii (Latham, 1790) Cuculus sonneratii Latham, 1790, Index Ornithologicus, 1, p. 215. (India ⫽ North Cachar Hills) Polytypic. Four subspecies. Cacomantis sonneratii sonneratii (Latham, 1790); Cacomantis sonneratii musi-

cus Ljungh, 1804; Cacomantis sonneratii fasciolatus Müller, 1843; Cacomantis sonneratii waiti Baker, 1919. Other names: Penthoceryx sonneratii (Latham).

Banded Bay Cuckoo Cacomantis sonneratii 431

Description ADULT: Sexes alike, above, head and back bright rufous or bay barred with dark brown, the rufous and dark bars equal in width (c. 2 mm), rump dark brown with rufous bars, wing coverts rufous barred dark brown, wing blackish brown finely notched rufous on outer webs of inner secondaries, tail graduated, rufous and dark brown, the central feathers T1 with rufous edge and dark brown center extending into bars that reach 60% to edge of feather, the outer feathers T2 to T5 more extensively rufous with dark brown bars, a subterminal black band and a terminal white tip, feathers with narrow tips; face with whitish eyebrow, white line under eye; underparts, throat to under tail coverts whitish with fine wavy brown bars, the dark bars 0.8-1 mm wide and the distance from bar to bar 4–6 mm, bend of wing white, underside of flight feathers of wing pale rufous, whitish at base, with a pale bar across inner primaries and secondaries, wing lining whitish with fine dark bars; iris yellow, bill black and broad, the base of lower bill greenish gray, feet gray. JUVENILE: Similar to adult in reddish-brown plumage, the crown and upper back with white tips to feathers, wing and tail flight feathers rufoustipped, face more whitish; underparts with bars narrow and irregular; feet olive-green. NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, MZB, ROM, RMNH, UMMZ, USNM,WFVZ, ZRC.

Subspecies Cacomantis sonneratii sonneratii (Latham, 1790); plumage above, bright rufous or bay barred with dark brown; India, Nepal, Bhutan, Bangladesh, Burma, Thailand, Laos and Vietnam, Malay Peninsula; Cacomantis sonneratii waiti Baker, 1919; plumage above darker brown; Sri Lanka; Cacomantis sonneratii fasciolatus Müller, 1843; smaller; Sumatra, Borneo and the Philippines (Palawan); Cacomantis sonneratii musicus Ljungh, 1804; smaller and plumage above paler brown; Java, Bali. Müller placed the locality for his Cuculus fasciolatus, in Temminck (1843), as “Java en Sumatra”. Robinson and Kloss (1923a) established the name for

Sumatra birds as fasciolatus, and later (Robinson and Kloss 1923b) noted that “Birds from Sumatra are darker than those from Java or from the Malay Peninsula”; that is, C. s. fasciolatus. Later, Junge (1948) proposed a new name for Sumatra birds, C. s. schlegeli, but this is a synonym of fasciolatus. Birds from the Malay Peninsula, described as C. s. malayanus by Chasen and Kloss (1931), were said to be more brightly colored than C. s. sonneratii. Before the Raffles Museum series was dispersed in part to other museums, Gibson-Hill (1949a) noted that Malay birds “do not appear to differ appreciably in color from some of the specimens of P. s. fasciolatus from Singapore and Sumatra”.

Measurements and weights C. s. sonneratii. India: Wing, M (n ⫽ 10) 119–125 (121.1 ⫾ 1.97), F (n ⫽ 7) 120–128 (123.7 ⫾ 3.0); tail, M 104–120 (113.6 ⫾ 4.9), F 105–119 (111.9 ⫾ 5.2); bill, M 19–21.4 (20.0 ⫾ 0.9), F 18.6–21.1 (19.6 ⫾ 0.9); tarsus, M 15.7–18.4 (17.2 ⫾ 0.9), F 15.1–19.3 (17.1 ⫾ 1.5) (AMNH, FMNH, UMMZ); C. s. waiti, Sri Lanka: Wing, M (n ⫽ 6) 119–126 (123 ⫾ 2.7), F (n ⫽ 3) 122–125 (123.5) (AMNH, BMNH); C. s. fasciolatus, Borneo and Sumatra: Wing, M (n ⫽ 6) 102–113 (107.7 ⫾ 4.1), F (n ⫽ 10) 103–113 (106.6 ⫾ 3.8) (AMNH, BMNH, RMNH,WFVZ); C. s. musicus, Java: Wing, M (n ⫽ 12) 105–116 (109.6 ⫾ 3.4), F (n ⫽ 10) 108–114 (110.8 ⫾ 2.1) (AMNH, RMNH, USNM). Weight, F (n ⫽ 2) 27.9–34.3 (31.1), U (n ⫽ 5) 32–38 (34.9) (Price 1979, Becking 1981, UMMZ, WFVZ). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 1 0 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 22 cm. Small rufous cuckoo with a white eyebrow streak, dark cheek, plumage with fine wavy bars (white eyebrow unique, bird more finely barred above and below than C. merulinus or C. variolosus) and bill broader across base.Tail differs from C. merulinus in the dark center of the rectrices (rufous bars continuous across shaft or nearly so in C. merulinus).

Voice The cadence song is a high-pitched whistled fournote “smoke-yer-pepper”, at 2.4 kHz and lasting

432 Banded Bay Cuckoo Cacomantis sonneratii 1 sec, each note drops in pitch and the notes in series drop in pitch; the song faster and higher in pitch than in Indian Cuckoo Cuculus micropterus. The cuckoo also has a rising call of 3–4 slow whistled notes on one pitch, then 3–6 faster notes rising in pitch then coming to a sudden stop,“fee, fee, fee, fee, fihihi-hihi fihihihi-hi-hi fihihi-hihi”, starting at 2 kHz and rising to 2.5 kHz and it lasts a few seconds (Smythies 1940, Ali and Ripley 1969, King and Dickinson 1975, White 1984, MacKinnon and Phillipps 1993, Kotagama and Fernando 1994, Scharringa 1999, Wells 1999, Sheldon et al. 2001, Supari 2003).

Davison and Fook 1995, 1996, Bishop 1999, Grimmett et al. 1999, Wells 1999, Robson 2000, Sheldon et al. 2001, Thomas and Poole 2003).

Range and status

Food

Indian subcontinent (except arid northwest), southern China (Huidong in southern Sichuan, Gengma and Xishuangbanna in Yunnan, Guilin in Guangxi), Burma, Thailand, Laos, Cambodia (once), E Tonkin, Vietnam, Sri Lanka, Sumatra (throughout, also Riau, Batu and Enggano islands), Java, Bali, Borneo and the Philippines (Palawan). In India they are most common in the south and the Western Ghats, and they occur in Orissa and the Deccan plateau, and north of the Gangetic Plain they are in the foothills of the Himalayas from Uttar Pradesh to Arunachal Pradesh and the southern Assam hills including Meghalaya, Cachar and Nagaland, Sylhet, and in NE Bangladesh. Seasonally migratory in China, partially migratory in India where they occur mainly in the monsoons, and resident elsewhere (Delacour and Jabouille 1931, Smythies 1940, 1981, van Marle and Voous 1988, Cheng 1991, Dickinson et al. 1991, Lekagul and Round 1991,

Insects, mainly caterpillars (Pieridae, Notodontidae, Arctiidae), bugs, grasshoppers, crickets and termites (Sody 1989, Sheldon et al. 2001).

?

?

Habitat and general habits Open broad-leafed forest, primary forest, forest edge, secondary forest and scrub, casuarinas, gardens, cocoa plantations and cultivated lands; lowlands to 900 m, rarely to 1500 m, in some areas as high as 2400 m in breeding season. Males sing from tall exposed trees. The cuckoos are foliage gleaners and they also make aerial sallies into termite emergences.

Breeding and life cycle In India season varies with that of its hosts: near Bombay the cuckoos lay from February to August, and in Assam April to August mainly in May and June (Baker 1934), in the Eastern Ghats they sing much of the year although most commonly from March to May (Price 1979), in Sri Lanka young are seen in June and October (Phillips 1948), in the Malay Peninsula the adults sing from January to May and fledglings are seen late May to late July (Medway and Wells 1976, Wells 1999), in Sumatra there was an egg in April (Voous 1951), in Java there are egg records from March to June, also in February and October (Kuroda 1936, Hoogerwerf 1949), and in Sabah laying females were taken in March (Sheldon et al. 2001). Brood-parasitic. Hosts are Common Iora Aegithina tiphia in India, Malay Peninsula and Java (Harrison 1969, Becking 1981, Robertson and Jackson 1992, Wells 1999); a reported “Plaintive Cuckoo C. merulinus” juvenile fed by Green Iora A. viridissima (Wilkinson et al. 1991) may be this cuckoo. Other host records include Red-whiskered Bulbul Pycnonotus jocosus in India (Robertson and Jackson, 1992), a juvenile fed by White-bellied Yuhinia Yuhinia zantholeuca in peninsular Thailand, and Scarlet Minivet Pericrocotus flammeus in Sri Lanka (Phillips, 1948), and perhaps other bulbuls and small babblers such as Stachyris (Baker 1906, 1908a, 1927,

Plaintive Cuckoo Cacomantis merulinus 433 1934, 1942, Ali 1996). Eggs are similar to eggs of iora, white with sparse reddish brown marks (Baker 1906, Abdulali 1943, Becking 1981); 18 ⫻ 14 mm in Java where an egg laid by a bird in the hand was whitish tinged greenish with markings of brown

and purplish (Hellebrekers and Hoogerwerf 1967, Becking 1981), and 19.1 ⫻ 14.5 mm in Sumatra (Voous, 1951). Incubation and nestling periods are unknown. The nestling evicts the host eggs and young from the nest (Becking, 1981).

Plaintive Cuckoo Cacomantis merulinus (Scopoli, 1786) Cuculus merulinus Scopoli, 1786, Delicae Florae et Faunae Insubricae, 2, p. 89. [⫽ Panay, Philippines] Polytypic. Four subspecies. Cacomantis merulinus merulinus (Scopoli, 1786); Cacomantis merulinus lanceolatus (S. Müller, 1843); Cacomantis merulinus querulus Cabanis and Heine, 1863; Cacomantis merulinus threnodes Cabanis and Heine, 1863.

Description ADULT: C. m. querulus, sexes alike (in part), above forehead and crown brownish gray, back brown, wing coverts brown, wing brown, rump brown with black streaks, tail blackish with white tips, inner tail feathers blackish with buff notches, outer web of outer tail feather completely barred white from below; underparts, throat and upper breast gray (sometimes washed rufous forward to chin), belly and under tail coverts rufous, bend of wing white, under wing coverts rufous; females (missexed specimens of querulus?) similar to males in one plumage phase; most females rufous plumage barred above and below, the crown and back dark brown with rufous bars, rump barred rufous and brown, tail dark brown with rufous bars complete (or nearly complete) across T1 to T5, throat and breast to belly and under tail coverts whitish with dark brown bars, under wing coverts whitish with dark brown bars; eye-ring blackish, iris red to brown, some with yellowish outer circle, bill blackish above, yellow below, feet yellow. JUVENILE: Above crown light rusty brown streaked darker brown, back and wing coverts barred rusty and dark brown, rump more streaked than barred, wing barred rusty and dark brown, tail blackish, central tail feathers with rufous notches nearly forming a bar, outer tail feathers barred rufous; underparts, throat whitish streaked black, breast

whitish barred black; iris brown, bill black, gape orange, legs greenish. Birds molt directly from juvenile to adult plumage. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, CM, CMNH, DMNH, FMNH, MCZ, MVZ, MZB, RMNH, SMF, UMMZ, USNM, UWBM,TISTR, YPM, ZMUC, ZRC, ZSM.

Subspecies Cacomantis merulinus querulus Cabanis and Heine, 1863; underparts rufous buff (female whitish with dark brown bars below); E India in eastern states from Assam and Bengal, and Bangladesh to S China, Hainan, Burma, northern and central Thailand, Laos and Vietnam; migrates south to eastern peninsular India; Cacomantis merulinus threnodes Cabanis and Heine, 1863; underparts paler than C. m. querulus and darker than C. m. merulinus, Malay Peninsula and Sumatra; Cacomantis merulinus lanceolatus (S. Müller, 1843); underparts grayish white; Java; Cacomantis merulinus merulinus (Scopoli, 1786); underparts gray to buff adult females unbarred, like the males; the Philippines, Borneo, and Sulawesi.

Systematics Gray-bellied Cuckoo C. passerinus is sometimes called conspecific with C. merulinus, but they differ in plumage and songs. The distributions are allopatric. Although Jerdon (1862:335) commented that “the two interbreed with each other . . . in Bengal you meet with every variation and shade of intermediateness”, intermediate specimens are rare. Hume (1888) examined 27 specimens from Assam,

434 Plaintive Cuckoo Cacomantis merulinus all in rufous plumage. In the duars of East Bengal and Assam and in the Naga, Lushai, Khasi and Garo hills, of 18 specimens taken by Koelz and Rup Chand from 1949 to 1955 and now in FMNH and UMMZ, 16 are rufous querulus, as are earlier Cachar and Assam (Changpu, 7000’) specimens in Tring. The westernmost localities reported for C. merulinus are in Nepal, Sikkim, Bhutan, Bihar, and Bhutan; but these either are not well documented or are wintering birds. A record in Nepal (Hodgson, BMNH 59.3.4.402, adult) has no date and no specific locality, and Hodgson’s Nepal localities have been questioned, Cocker and Inskipp 1988). A record in Bihar (C. M. Inglis, Darbhanga, 7 November 1903,YPM 42593), in adult plumage with retained juvenile secondaries and nape feathers, by season was a wintering bird; Cacomantis taken there in June to August are gray (YPM 42590, 42591; Inglis et al. 1920, Inglis 1923). Birds in Raipur (Hume collection, January and February 1871), were wintering; others Cacomantis there are gray. Rufous birds in the duars, the Himalayan foothills near Bhutan, were an adult female in January (BMNH 87.12.2.424, Mandalli coll.) and a juvenile, no date (BMNH 79.11.28.648, Pemberton coll.), by date and collector were in India as now recognized, as the political boundary changed and the duars are now in western Assam (Rasmussen and Prys-Jones 2003). Inskipp and Inskipp (1985) listed sight records of rufous birds in Nepal (Chitwan on 21 February 1986, Kosi Barrage on 9 February 1989) in lowland terai both in the nonbreeding season. Biswas (1960) noted no rufous birds in Nepal, and only gray birds were noted in Dudwa NP, India, on the Nepal border ( Javed and Rahmani 1998) and west at Dehra Dun (Mohan 1997). Ali (1962) noted none in Sikkim. In the Eastern Ghats a rufous bird was taken at Cumbum, Andhra Pradesh (BMNH 1949.5.1197 ⫽ 1949.Whi.1.16681, 26 Dec 1929, “adult plumage except for some remains of the juvenile dress” (Whistler and Kinnear 1934:521. No C. m. querulus in this western region were breeding season adults. There may be a narrow zone of sympatry with limited interbreeding between C. m. querulus and C. passerinus. A bird (UMMZ 142889) taken by Koelz at Agia, Assam, on 2 May 1950 is darker gray on the back than the Assam series of C. m. querulus,

gray from throat to the upper belly and flanks, and rufous on the lower belly and under tail coverts, an intermediate between C. passerinus and C. m. querulus. Koelz noted on the label “calling clearly C. passerinus notes”. Biswas (1951) reported both forms taken by Koelz at Barapani in the Khasi Hills (adult male querulus on 23 May 1949, FMNH 229725; adult male passerinus on 14 June 1949, FMNH 229726). Biswas reported that Koelz found C. m. querulus above 3000 feet and C. passerinus at lower altitudes, and in India generally querulus occurs at higher altitudes than passerinus. In Kaziranga NP, Assam, at 93° E, merulinus is a common local migrant; passerinus is uncommon and not known to breed (Barua and Sharma 1999). Inglis et al. (1920) and Ali and Ripley (1969) listed both passerinus and querulus for Jalpaighuri district in Bengal (88°30' E to 90° E) and in Bhutan, Sikkim and Nepal, but their regions of reference extended across the area of both forms.Whistler (1935, 1949) noted gray passerinus east to the Brahmaputra River; and rufous querulus in Assam, Eastern Bengal and occasionally farther west. A few female cuckoos are intermediate in plumage between rufous and gray: (1) Phulbari, Garo Hills, taken 20 April 1950 (UMMZ 142877) has crown and back gray indistinctly barred rufous, breast barred rufous and belly whitish indistinctly barred gray, (2) Darjeeling, taken 8 August 1945 by W. H. Mathews (YPM 42591) has a mix of rufous and gray plumage, (3) Sikkim, no date, taken by H. Whitely (FMNH 301950) is in gray plumage with rufous feathers growing on nape and back, and (4) Bichia, Madhya Pradesh, taken 7 July 1946 by Koelz (FMNH 229716) west of the breeding range of C. m. querulus, has a mixed rufous and gray plumage. These may represent interbreeding between rufous and gray birds. In Burma, Smythies (1940) described rufousbellied C. m. querulus in the text, but added graybellied passerinus in the appendix from a list of birds collected by Stanford in the Upper Chindwin and northeast Burma. Later he noted “belly not always red as shown in the plate” and listed both passerinus and querulus (Smythies 1953: 598), passerinus in the Upper Chindwin and northeast Burma to 6000’, which may refer to birds in Yunnan and querulus in the other areas of Burma. Stanford reported

Plaintive Cuckoo Cacomantis merulinus 435 passerinus at Myitkyina, Lawsun and Namyin Valley in northern Burma and Tenggueh in Yunnan (Stanford and Ticehurst 1939). However, Ticehurst described the throat and breast as gray and the belly “ferruginous”, the color of rufous querulus not gray passerinus, while the Tenggueh birds were rufous, and Rothschild (1921, 1925) identified them as querulus. Earlier reports on birds of northern Burma described rufous-bellied querulus on the Chindwin (Mears and Oates 1907), as did a later Stanford collection in the Myitkyina area (Stanford and Mayr 1941). All specimens from Burma and Yunnan are rufous C. m. querulus not gray C. passerinus.

Measurements and weights C. m. querulus, Burma, Thailand, Laos, Vietnam, China (Yunnan and Fujian): Wing, M (n ⫽ 34) 98–114 (110.1 ⫾ 5.1), F (n ⫽ 11) 106–117 (110.1 ⫾ 5.2); tail, M 94–122 (108.6 ⫾ 7.4), F 95–114 (105.3 ⫾ 6.7); bill, M 16.5–19.2 (17.1 ⫾ 0.8), F 16.2–17.9 (1.1 ⫾ 0.6); tarsus, M 14.2–17.9 (15.7 ⫾ 1.2), F 12.9–18.6 (15.4 ⫾ 1.5) (AMNH, ANSP, FMNH, UMMZ, USNM, ZMUC); C. m. threnodes, Malacca, Sumatra and Borneo, wing, M (n ⫽ 6) 96–104 (99.8 ⫾ 2.9), F (n ⫽ 7) 95–199 (97.2 ⫾ 2.4) (FMNH, MZB); C. m. lanceolatus, Java: wing, M (n ⫽ 6) 96–107.5 (103.9 ⫾ 4.0), F (n ⫽ 7) 98–109 (101.6 ⫾ 4.4) (MZB, ZRC); C. m. merulinus, the Philippines: Wing, M (n ⫽ 18) 97–118 (103.9 ⫾ 4.1), F (n ⫽ 11) 101–112 (105.2 ⫾ 3.6) (FMNH, ZMUC). Weight, Malay Peninsula: U (n ⫽ 4) 21.3–26.6 (Wells 1999); Sumatra, F (n ⫽ 1) 25 (MZB); Borneo: M (n ⫽ 1) 26, F (n ⫽ 1) 27.0 (ZMC,Thompson 1966), the Philippines: M (n ⫽ 1) 26.6, F (n ⫽ 2) 22.4–27.2 (24.8) (FMNH). Wing formula, P8 ⬎ 9 ⱖ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 1 0 ⬎ 2 ⬎ 1.

Field characters Overall length 22 cm. Small cuckoo with brown upperparts, gray throat and upper breast.Adult overlaps in range with Brush Cuckoo C. variolosus and differs in gray throat and breast (rusty in C. variolosus); outer tail feathers barred white below (notched white in C. variolosus). Barred phase of female is much like in rufous phase Gray-bellied Cuckoo

C. passerinus but is distinctly barred; and is like Banded Bay Cuckoo C. sonneratii, which has narrow tips to the tail feathers and a broad bill; C. merulinus differs by lack of white streak over eye, and tail with tan to rufous bars continuous across feathers in barred female phase (rufous restricted to V-shaped notches on edges of feathers in C. variolosus and C. sonneratii). Juvenile, plumage barred above and below, crown buff to rufous streaked with black (barred in C. variolosus), throat streaked (barred in C. variolosus), rump streaked or barred with the rufous continuous (barred with the dark brown continuous across the feathers in C. variolosus), back with pale bars wider than dark bars. In flight, has a white band at base of the flight feathers.

Voice Sings day and night in breeding season. (1) Cadence song is an accelerating trill on a descending scale, “tee-tee-tee-tee-tita-tita-tita-tita-tee”, 3–4 slow notes then 4 or more rapid notes descending the scale, heard in the breeding season (Smythies 1940). (2) Rising song (or ascending song) is a shrill threenote phrase “tay-ta-tee”, “please-don’t-do-it” or “eat more froueet” repeated on a rising scale, with each phrase higher in pitch than the one before it (Figure 6.4) (Livesey 1938c, Coomans de Ruiter and Maurenbrecher 1948, Coomans de Ruiter 1950, Smythies 1953, 1960, 1981, 1986, King and Dickinson 1975, Fleming et al. 1979,Watling 1983, White 1984, Hails 1987, Gonzales and Rees 1988, Smith 1993a, Holmes and Phillipps 1996, Scharringa 1999, Wells 1999, NSA, RBP). Blyth (in Jerdon 1862) thought calls of rufous-bellied birds in Upper Burma were less plaintive than Indian C. passerinus, and Davison (in Hume and Davison, 1878) thought the same about calls in southern Burma.The songs and calls of C. merulinus appear to be the same in Assam, Burma, Malay Peninsula, Java, Sulawesi and Borneo.

Range and status Southern and southeast Asia to the Greater Sunda Islands, Sulawesi and the Philippines. In southern China they occur from southern Tibet and Yunnan to Guangxi, Fujian and Hainan, seasonal from February to October (Cheng 1991,Vuilleumier 1993). In India they are in Assam. In southeast Asia they

436 Plaintive Cuckoo Cacomantis merulinus other soft-bodied insects including aphids; and fruit (Legge 1880, Ali and Ripley 1969, Sody 1989).

Displays and breeding behavior Males call from regular sites, territorial, often concealed in foliage as they sing. Courtship feeding, male fed a caterpillar to a female (Wells 1999).

Breeding and life cycle occur in Bangladesh, Burma,Thailand, Laos,Vietnam and the Malay Peninsula, and in Sumatra, Borneo and Sulawesi, and perhaps in Togian Islands. In the Philippines they occur on Balbac, Basilan, Bohol, Cebu, Leyte, Luzon, Masbate, Mindanao, Mindoro, Palawan, Panay and Siargao, and on Malamaui, Rasa and Biliran; other records are misidentifications of Brush Cuckoo C. variolosus (Dickinson et al. 1991). Migratory in China, resident in most of their range. In Assam they are mainly resident (Stevens 1915) and local migrants at Kaziranga NP (Barua and Sharma 1999); a few appear further west in winter. In Thailand they are common in the north in summer and in the south in winter (Legge 1880, Hartert 1925a, Riley 1938, Stresemann 1940, White 1944, Phillips 1948, Henry 1971, White and Bruce 1986, Kotagama and Fernando 1994, Robson 2000a).

Habitat and general habits Open woodland, secondary forest, scrub-covered hillsides, brush, bamboo, gardens, cultivated areas, to towns and villages, grassy plains and swamps (Stanford 1931, Ali and Ripley 1969, Becking 1981, Smythies 1986, Gonzales and Rees 1988, van Marle and Voous 1988, Robson 2000a). In Assam they are generally below 1000 m (Hume 1888); in Thailand seen to 1830 m (Robson 2000a). In Kalimantan, Sabah and Sulawesi they are in wooded lowlands (Coates and Bishop 1997, Holmes 1997, Sheldon et al. 2001). Plaintive Cuckoos are arboreal, active and restless in foliage canopy; and occur more in the lowlands than in the hills.

Food Insects, mainly caterpillars (hairy Saturniidae, hairless Notodontidae), beetles, bugs, termite soldiers,

In eastern India in Assam they breed from April to August (Baker 1934). Late breeding at Takubama, Naga Hills, is indicated by a short-tailed, rufous fledgling (UMMZ 142883) taken by Koelz on 27 Sept 1950. In Burma they breed from April to July (Osmaston 1916, Smythies 1986), in northern Thailand from April to August (Lekagul and Round 1991), in Malay Peninsula fledglings are seen from early June to late August (Wells 1999), in Java eggs occur from April to July and in October (Hoogerwerf 1949), in Borneo they breed in March, May and June (Nash and Nash 1988, Sheldon et al. 2001) and in Sulawesi a young bird was out of the nest in September (Riley 1924). Brood-parasitic. Hosts are mainly warblers with concealed nests, either globe-like nests with small side entrances or deep purselike nests with slit-like entrances: Yellow-bellied Prinia P. flaviventris and Black-necked Tailorbird O. atrogularis in Malay Peninsula (Wells 1999), P. flaviventris, C. juncidis and Olive-backed Tailorbird O. sepium in Java (Hellebrekers and Hoogerwerf 1967, Becking 1981), prinias and tailorbirds in Borneo (Smythies 1957, 1999) and Grey-backed Tailorbird O. derbianus in the Philippines (Hornskov 1995). In Sabah, the cuckoo is said to parasitize Common Iora Aegithina tiphia, Streaky-breasted Spiderhunter Arachnotherea affinis and Yellow-bellied Prinia (Sheldon et al. 2001). A report of a juvenile cuckoo fed by Green Iora A. viridissima (Wilkinson et al. 1991) may be of a Banded Bay Cuckoo C. sonneratii. Eggs are dull, light buff to green with light olive marks like the eggs of the host species, with three distinct variations described in Java; the shape oblong to ovate, 20 ⫻ 14 mm (Baker 1906, 1907, 1934, Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967, Becking 1981). Incubation and nestling periods are unknown.

Grey-bellied Cuckoo Cacomantis passerinus 437

Grey-bellied Cuckoo Cacomantis passerinus (Vahl, 1797) Cuculus passerinus Vahl, 1797, Skrivter af Naturhistorie-Selskabet, Kjøbenhavn , 4, Heft 1, p. 57. (Tranquebar [⫽ India]) Monotypic. Other common names: Indian Plaintive Cuckoo.

Description ADULT: Sexes differ, at least in part. Male, above the head and back dark gray lightly glossed green, wing coverts dark gray, wing dark gray, tail blackish with white tips and outer rectrices barred white; underparts dark gray, whitish belly (sometimes) and under tail coverts, bend of wing white, under wing coverts barred white. Female, a few (mis-sexed?) are gray phase like male, belly lightly barred whitish, tail barred on more than outer rectrices, feathers with white edge; most females are rufous phase, variable in appearance, above the crown rufous, sometimes streaked with dark brown, nape rufous, back bright rufous barred with dark gray-brown, the rusty bars wider than the dark bars in some and about the same width in the darker females), rump rufous unbarred or nearly so, tail rufous, the outer rectrices with a blackish subterminal bar, face rufous; below throat and upper breast rufous with indistinct barring on lower breast, the intensity and extent of the rufous varies among females, belly and under tail coverts white with blackish bars, bend of wing white, under wing coverts barred white; eyering gray, iris red with a yellow outer ring to reddish brown, bill brown to black, feet yellow-olive. JUVENILE: Plumage variable, generally noted by gray upperparts with variable indistinct marks of rufous, and below by indistinct rufous markings on breast on a mostly gray or white-and-gray barred belly. Above, the crown and back dark gray with some indistinct rufous bars, rump and upper tail coverts unbarred or the feathers tipped rufous, tail blackish with rufous notches on the vanes of the inner rectrices and white notches on the outer rectrices, face dark with variable gray or rufous blotches, throat and upper breast whitish or rufous and barred gray, fine whitish bars on belly and with a dark gray tail with some white bars; others (most

or all of these are females) are rufous with crown light rusty brown streaked darker brown, back and wing barred rusty and dark brown (rusty bars as wide as dark brown bars), rump rufous with some dark bars, tail with rufous edge or without rufous, the face rufous with indistinct blackish marks; below the throat and breast whitish with a diffuse rusty wash, lower breast, belly and under tail coverts whitish barred gray, some birds mostly whitish and others gray with fine vermiculations, bend of wing white, under wing coverts barred brownish and white; iris blackish, bill dark brown, feet light yellow. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, FMNH, FU, MCZ, RMNH, ROM, UMMZ, USNM, YPM, ZMUC, ZSM.

Systematics Gray-bellied Cuckoo C. passerinus is sometimes recognized as a subspecies of C. merulinus, and their host species are the same. Both forms occur in India and they are mainly allopatric. In northern Pakistan to Abbottabad, in the northern hill country of India from Darjeeling west to Kumaon, Simla, Mussoorie, Dharmasala, Kangra, Kulu Valley and Sikkim, and Nepal, all specimens are gray-bellied passerinus, as are most birds in peninsular India (116 of 120 adults in museum collections), whereas birds in upper Assam and Bangladesh are rufous-bellied querulus (Stevens 1915, 1925, Ali and Whistler 1937, Biswas 1960, Ali and Ripley 1969, Inskipp and Inskipp 1985, Pandey et al. 1994, Inskipp et al. 2000; BMNH).

Measurements India and Nepal: Wing, M (n ⫽ 18) 105–120 (114.6 ⫾ 4.5), F (n ⫽ 5) 110–115 (113.0 ⫾ 2.7); tail, M 98–115 (104.2 ⫾ 5.1), F 87–115 (100.0 ⫾ 5.2); bill, M 15.3–17.7 (16.6 ⫾ 0.8), F 15.9–18.3 (18.1); tarsus, M 13.3–17.7 (15.4 ⫾ 1.5), F 15.6–18.0 (17.1 ⫾ 2.2) (AMNH, FMNH,YPM, ZMUC).

438 Grey-bellied Cuckoo Cacomantis passerinus Wing formula, P8 ⬎ 9 ⱖ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 22 cm. Small gray cuckoo distinguished by the lack of streaks or barring; rufous female differs from rufous-phase Plaintive Cuckoo C. merulinus in being less barred and in having the contrasting rufous head. Juveniles differ from C. merulinus in having a dark gray crown with irregular rufous blotches (the crown not pale rufous with dark streaks). In Sri Lanka, the bright uniformly rufous birds in winter are female C. passerinus rather than C. merulinus. No white streak over the eye (white streak present in Banded Bay Cuckoo C. sonneratii).

Voice Sings day and night in breeding season. A plaintive whistle “ka.teer” repeated at 1-sec intervals (Fleming et al. 1979, Martens and Eck 1995), the “ka-veer” of Jerdon (1862), “whe-whew” of Oates and Blanford (1895), “how few” of Frome (1946), and “piteer” or “p’teer” or “peter-peter” of Ali (1962, 1996) and Price (1979). Another call is a descending whistled series “pee-pipee-pee-pipeepee” (Ali 1962; Figure 6.4).

Range and status Hills of northern Pakistan (Hazara to Murree Hills), northern India from base of western Himalayas to Nepal and Sikkim, and south from the Gangetic Plain through most of peninsular India, though not in the drier northwest. In India it is common except in the drier parts and in montane regions, and is most numerous on Deccan Plateau (Baker 1934) south to Nilgiri hills and Wynaad, Kerala, also seasonally in the Eastern Ghats (Price 1979). In Nepal it is an uncommon summer visitor. In Bangladesh it occurs locally in summer. In Sri Lanka, Lakshadweep (Laccadive Is) and Maldive Is it arrives with the dry northeast winds in October, remains over the winter, and departs in April; a winter visitor, it does not sing and there are no breeding records in Sri Lanka (Legge 1880, Hartert 1925a, Baker 1934, Stresemann 1940, White 1944, Phillips 1948, Henry 1971, Inskipp

and Inskipp 1985, Roberts 1991, Kotagama and Fernando 1994). Occasional in winter in Oman, 7 Nov 1989 (Eriksen 1991). Seasonally migratory in the north as far south as Orissa, resident in the south.

Habitat and general habits Open woodland, secondary forest, scrub-covered hillsides, brush, gardens, cultivated areas, to towns and villages, grassy plains and swamps (Ali and Ripley 1969, Becking 1981, Zacharias and Gaston 1983). In Assam they occur below 1000 m (Hume 1888). In Nepal they are rare as high as 1400–2100 m in nearly treeless areas and in farmlands with trees (Martens and Eck 1995). In Sri Lanka in winter they are common in dry zones and in gardens more in lowlands than in the hills.Arboreal, active and restless in foliage canopy.

Food Insects, mainly caterpillars (Legge 1880, Ali and Ripley 1969).

Displays and breeding behavior Males call from regular sites, territorial, concealed in foliage as they sing.

Breeding and life cycle In Pakistan they breed from March to September (Roberts 1991), in eastern India they call in Assam from April to August, in the Eastern Ghats they call between April and June, in southern India in the Nilgiris they call from August to October and at

Brush Cuckoo Cacomantis variolosus 439 Kotagiri a feathered nestling cuckoo was taken in a prinia nest in December (Baker 1934, Price 1979, BMNH). Brood-parasitic. Hosts are mainly warblers with concealed nests, either domed nests with small side entrances or deep purselike nests with slit-like entrances: Ashy Wren-warbler Prinia socialis, Fantailed Warbler Cisticola juncidis, Common Tailorbird Orthotomus sutorius and sunbirds (Nectarinia asiatica, N. zeylonica) in India (Baker 1906, 1908a, 1934,

Stevens 1925, Becking 1981, Bharucha 1982; BMNH). Eggs are dull, light buff to green with light olive marks like the eggs of the host species, oblong to ovate, 20 ⫻ 14 mm. Eggs of C. passerinus are the same size and color as eggs of C. merulinus in southeast Asia from Bengal,Assam and Burma to Malay Peninsula. Incubation and nestling periods are unknown. Tailorbirds desert 20% of their nests that have a cuckoo egg.The nestling evicts the host eggs from the nest (Legge 1880).

Brush Cuckoo Cacomantis variolosus (Vigors and Horsfield, 1827) Cuculus variolosus Vigors and Horsfield, 1827, Transactions of the Linnean Society, London, 1, part 1, p. 300. (Paramatta, New South Wales) Other common names: Square-tailed Cuckoo, Grey-breasted Brush Cuckoo, Rusty-breasted Cuckoo, Rusty-breasted Brush Cuckoo, Indonesian Cuckoo, Moluccan Brush Cuckoo, Greyheaded Cuckoo. Polytypic.Ten subspecies. Cacomantis variolosus variolosus (Vigors and Horsfield, 1827); Cacomantis variolosus sepulcralis (S. Müller, 1843); Cacomantis variolosus infaustus Cabanis and Heine, 1863; Cacomantis variolosus virescens (Brüggemann, 1876); Cacomantis variolosus aeruginosus Salvadori, 1878; Cacomantis variolosus websteri, Hartert, 1898; Cacomantis variolosus addendus Rothschild and Hartert, 1901; Cacomantis variolosus blandus Rothschild and Hartert, 1914; Cacomantis variolosus macrocercus, Stresemann, 1921; Cacomantis variolosus oreophilus Hartert, 1925.

Description ADULT: (Cacomantis variolosus sepulcralis) Sexes alike (in part), head gray, back brownish gray, wing coverts brownish gray, wing brownish gray, rump and under tail coverts dark (black as extensive as rufous) tail rufous with black bars (rufous does not extend across the feather), tips white; underparts, throat gray (some birds with rufous tinge, other birds with upper breast gray), breast and belly to under tail coverts rusty, no sharp demarcation of gray cheeks and rusty underparts, unbarred (male) to finely barred gray (female), tail below broadly barred brown with thin whitish notching on inner web, tip of rectrices white, bend

of wing white, under wing coverts darker rufous to buff than breast and belly, under wing slate with a broad white band across primaries and secondaries. Other adult females have a barred plumage, above barred dark brown and rufous (rufous bars as wide as dark bars), underparts from chin to under tail coverts whitish with even blackish bars 1 mm wide, the bars more widely spaced on under tail coverts, tail barred rufous on T3-5, under wing coverts whitish with narrow blackish bars. A few females in intermediate plumage are partly rufous and partly barred below, under wing coverts either rufous or whitish with narrow blackish bars or mixed, and no molt. Other females in mixed plumage are in molt with a few sheathed and growing unbarred rufous feathers, indicating a variable plumage that changes from barred to unbarred rufous.All these adults have dark tips to the primaries, not pale as in juveniles. In a sample from the Philippines, all males are rufous, whereas for females 27 are rufous, 13 are barred, and 8 are in mixed plumage (DMNH, FMNH, UMMZ, USNM). Eye-ring yellow, iris deep red brown or with inner ring rufous and outer ring whitish (male) to brown or light brown (female), bill black with yellow base below, inside of mouth in males orange or red, in females yellow or orange, feet yellow to gray-pink. JUVENILE: Head, back and wing barred rufous to yellowish buff and dark brown (dark bars wider than buff bars), the bars irregular in width, tips of flight feathers buff, tail dark brown, T1 to T3 all dark with rufous notches on inner and outer vane,

440 Brush Cuckoo Cacomantis variolosus T4 and T5 with bars white or rufous, pale bars complete or nearly complete across tail feathers T4 and T5; face dark brown irregularly marked with rufous buff, underparts white with blackish bars from breast to under tail coverts, the bars irregular or splotchy, bend of wing white, under wing coverts whitish with narrow black bars; eye-ring light gray, iris dark brown, bill black, feet dull yellow. Birds molt directly from barred juvenile plumage into unbarred adult plumage. NESTLING: C. v. variolosus naked at hatching, skin black, inside of mouth orange (AMNH 626018) to papaya red (UWBM 58699; Strahan 1994). SOURCES: AMNH, ANSP, BMNH, CM, CMNH, DMNH, FMNH, JFBM, MCZ, MVZ, MZB, RMNH, ROM, SMF, SMTD, TISTR, UMMZ, UWBM, WAM, YPM, ZMB, ZMUC, ZRC.

History and subspecies Geographic variation involves two sets of populations, the northern sepulcralis and the southern variolosus groups. The northern birds are smaller and more rufous, southern birds average larger and are more gray. The two sets are sometimes considered different species (e.g. Stresemann 1912, 1914, 1931); however, the songs are similar in northern and southern birds, the size and plumage color intergrade in the Moluccas, and the populations do not consistently differ in eye-ring color, with both gray and yellow eye-rings occurring in New Guinea and Australia, the feature often noted in field guides as distinguishing the two sets; in fact, yellow eye-rings occur throughout the range, while gray eye-rings are reported in adults only in southern populations, and gray eye-rings also occur in juveniles in northern populations.These northern and southern birds are genetically similar. Cacomantis variolosus sepulcralis (S. Müller, 1843); above brown, underparts rufous in males and some females, other females barred above and underparts whitish barred brown; widespread from S Burma (Tenasserim), Thailand, Malay Peninsula, Sumatra, Java, Borneo and Lesser Sundas (Lombok, Sumbawa, Flores, Sumba) to the Philippines;

Cacomantis variolosus virescens (Brüggemann, 1876); above darker greenish brown, underparts darker rufous; Sulawesi and neighboring islands (Butung,Tukangbesi, Banggai, Peleng); Cacomantis variolosus aeruginosus Salvadori, 1878; plumage darker than virescens; C Moluccas (Sula Islands, Buru, Ambon, Seram); Cacomantis variolosus infaustus Cabanis and Heine, 1863 (includes major Salvadori 1880, fortior Rothschild and Hartert 1914, oblitus Hartert 1925, heinrichi Stresemann 1931, chivae Mayr and Meyer de Schauensee 1939, and obiensis Jany 1955); above darker brownish gray with no gloss, breast gray, belly rufous gray; eye-ring whitish (both “yellow” and “gray” specimen label notations in this population), bill slender (thicker in oreophilus); juvenile darker, tail without pale bars or pale bars only below on outer feathers, under wing coverts dark buff; N and S Moluccas (Morotai, Tidore, Ternate, Halmahera, Bacan, Obi, Seram Laut, Watubela, Kai); western Papuan islands (Gebe,Waigeu, Salawati, Kofiau, Misool, Goram), Biak, NW and N New Guinea east to Sepik River and islands northeast of New Guinea as far as Umboi; Cacomantis variolosus oreophilus Hartert, 1925; overlap in plumage color but wing smaller and bill larger than infaustus, plumage of birds in highlands darker, above more brownish olive with dull greenish gloss, adult females are either unbarred (most birds) or barred rufous, and underparts variably barred, darker than females in Australia; E and S New Guinea highlands; Cacomantis variolosus blandus Rothschild and Hartert, 1914; back gray, throat, upper breast and under tail coverts gray, belly rufous; Bismarck Archipelago (Admiralty Islands); Cacomantis variolosus websteri Hartert, 1898; underparts slaty gray, tail short; Bismarck Archipelago (New Hanover); Cacomantis variolosus macrocercus Stresemann, 1921; tail long, belly slaty gray, rufous or intermediate, eye-ring gray or yellow; Bismarck Archipelago (except Admiralty Islands and New Hanover); Cacomantis variolosus addendus Rothschild and Hartert, 1901; blackish above, throat gray, breast and belly rufous buff, eye-ring yellow; Solomon Islands;

Brush Cuckoo Cacomantis variolosus 441 Cacomantis variolosus variolosus (Vigors and Horsfield, 1826); male, head light gray, back unmarked brownish gray, underparts with throat and upper breast pale gray, breast and belly buff; eye-ring yellow or more often gray, iris dark brown; female, two color phases, an unbarred phase similar to adult male but gray of breast extends to belly and flanks, and upper secondary coverts finely spotted rufous to buff, under wing coverts light rufous; and barred phase with head and back to rump brown barred rufous (gray bars 2–3 mm, rufous bars 1.8–2.2 mm), underparts variable with barring gray or brown, the bars even in width, under wing coverts whitish with narrow bars (female plumages are variable, and intermediates occur between the two extremes); juvenile, the head and back barred brown, wing and tail feathers with buff edge notches, underparts white with fine irregular brown bars, eye-ring dark gray, narrower than in adult; northern and eastern Australia and Timor, winters to Aru Islands, the Moluccas, Kai Islands, New Guinea and western Papuan islands, and one record in Taliabu. In the Philippines, earlier work (McGregor 1909) did not distinguish between C. merulinus and C. variolosus; the species were first distinguished by Hartert (1925): Rand (1951), Dickinson and Heucke (1986) and Dickinson et al. (1991) give additional history. C. v. everetti (Hartert 1925) from the Sulu Islands was based on two birds with more rufous on the throat, but the population is not distinct and the amount of rufous on the throat varies within other populations of Philippine C. v. sepulcralis. In the northern Moluccas, rufous birds known as “C. heinrichi” (Stresemann 1931) were once recognized as a species. In contrast to the original description, rufous birds in this region are not all smaller than gray birds. Birds identified as “heinrichi” are dark olive brown above, the crown more gray than back, throat is dark gray, breast and belly to under tail coverts are dark rufous, bend of wing white, under wing coverts are unbarred buff; eyering is yellow, iris dark brown with lighter outer ring, feet yellow. Juvenile “heinrichi” (AMNH 625805, 625810, 467572, and others with a mix of juvenile and adult plumage) are dark above, the remiges, greater upper coverts and tail feathers are barred rufous on outer web, and are barred on

the face, throat and breast. The main trait said to distinguish these birds from (other) C. variolosus in the northern Moluccas was the rufous underparts in both males and females. However, the plumages of both rufous and gray birds vary and the extreme forms intergrade in the northern Moluccas. The intergradation is probably due to repeated immigration from other populations and interbreeding, with rufous birds in the sepulcralis complex from the southern Moluccas and gray birds in the variolosus complex from New Guinea, much as the variable birds on the Lihir Islands in the Bismarck Archipelago resemble those of neighboring island groups. Hartert (1925a) considered the rufous birds of Halmahera to be a color phase of C. variolosus. Stresemann (1931) reported them as smaller than the sympatric C. variolosus infaustus, and he described them as a species C. heinrichi on the basis of size and plumage color. Specimens from the northern Moluccas (excluding Ternate trade skins, which may have come from elsewhere, Salvadori 1878a, van Bemmel 1948) vary in size; no consistent difference is apparent in plumage color between large and small birds, or in size of the more rufous birds and the gray birds, and the distribution of male wing lengths is not bimodal. As described below under “voice”, the birds appear to have similar songs and they are known by local peoples as the same species. In addition, certain C. variolosus in western New Guinea are similar in size and plumage to birds in the AMNH type series of heinrichi (YPM 74836), an adult taken at Wanvena with wing 121, is as small and nearly identical to these birds in plumage but has lighter rufous under tail coverts; KU 43513 from Geelvink Bay (⫽ Teluk Cenderawasih), a laying female, wing 114, is as dark as birds in the Moluccas, with the back dark brownish gray, underparts barred rufous and dark brown, under tail coverts rufous; and SMTD C-1869 from Kordo (Corrido, Biak) is small, dark and rufous.The genetic distance of “heinrichi” from other Cacomantis variolosus is less than that between other forms of C. variolosus (Figure 5.6). C. v. stresemanni Hartert 1925 on Seram was described as having a large bill, but the type series was only two specimens and the bills are not larger than all others in the Moluccas.

442 Brush Cuckoo Cacomantis variolosus Finally, the form C. v. obiensis Jany 1955 (the holotype male MZB 21,501 was originally identified as “heinrichi”) was said to be darker in plumage than birds elsewhere in the northern Moluccas; however, the type series (two males, two females) in MZB as well as two other MZB specimens from Obi fall within the range of variation of birds in the rest of the northern Moluccas, and obiensis is a synonym of C. v. infaustus. Another Moluccan form that has been recognized is major Salvadori 1880, described as a variety (“var.”) of Cacomantis assimilis (⫽ C. variolosus). Mees (1965, 1972) recognized major as having priority over C. v. oblitus Hartert 1925, also from the northern Moluccas (Batjan ⫽ Bacan). By the current Code (Article 45.6.4,1) (ICZN 1999), the name major has priority, particularly because of its recognition by Mees. Salvadori’s type series included seven specimens from Ternate and Halmahera; he also recognized the form from other islands: Mafor (⫽ Numfor), Duke of York and New Britain. Because birds in the northern Moluccas do not differ consistently from birds in the western Papuan region, the forms major Salvadori 1880 and oblitus Hartert 1925 are synonyms of infaustus Cabanis and Heine 1863. Plumage within the southern C. variolosus complex varies geographically, and many subspecies have been described from variation in length of wing and tail, color of upperparts and extent of rufous on the underparts (Hartert 1930). In the New Guinea region the plumage is variable (Mayr and Rand 1937, Rand and Gilliard 1947), and additional forms have been described, including C. v. fortior Rothschild and Hartert 1914 on islands near northeast New Guinea (in a small sample they were larger than infaustus (Rothschild and Hartert 1907, 1914b), while larger samples show overlap in measurements and bill width (Diamond and LeCroy 1979; AMNH)), C. v. chivae Mayr and Meyer de Schauensee 1940 on Biak (plumage darker, and birds smaller than infaustus), C. v. obscuratus Stresemann and Paludan 1932 on Numfor (plumage much like infaustus, some are darker than other New Guinea) (Rand and Gilliard 1967, Diamond 1972). Females have two color phases, gray and rufous; and C. v. infaustus has a more slender bill than C. v. oreophilus; the overlap in meas-

urements and width of bill prevent recognition of fortior. The bill of birds in southeast New Guinea is wider and thicker at the base, with variable length; these are oreophilus. In the Bismarck Archipelago, several subspecies are known, and these tend to be darker in plumage and some are longer-tailed than birds on New Guinea. In addition to the subspecies recognized (C. v. blandus, C. v. macrocercus, C. v. websteri), two specimens described as C. v. tabarensis Amadon 1942, from Tabar I, north of New Ireland, Bismarck Archipelago, have a large bill; other measurements and plumage are like that of C. v. macrocercus. On Lihir Islands north of New Ireland and east of Tabar I, birds are variable in plumage, the population perhaps an interbreeding mix of C. v. macrocercus and C. v. websteri. In the Solomon Islands, adult females occur in two color phases, and the barred phase is known to breed (Galbraith and Galbraith 1962); absent on Rennell Island (Mayr and Diamond 2001). Birds in western and northern Australia are smaller on average than birds in SE Australia but they overlap in size. The holotype of tymbonomus S. Müller 1843 (RMNH 88156) from Timor has wing 125, tail 111. Timor birds named by Bruce (in White and Bruce 1986) as C. v. whitei, from a series in AMNH, described with “narrow rufous edges to the grayish brown back . . .”, are in partial or complete juvenile plumage and look like birds in Australia. Because they were taken from January to April during the Australian breeding season, they appear to be residents and not wintering birds from Australia. The holotype of dumetorum Gould 1845 (ANSP 20028) from Port Essington, Australia, has wing 123 and the same plumage. Insofar as the plumage differences are accounted for by age, the birds in Australia are variable in plumage, and the sizes overlap, the birds from Timor, western, northern and eastern Australia are all considered C. v. variolosus; while C. v. dumetorum (Gould 1845), C. v. tymbonomus (S. Müller 1843) and C. v. whitei Bruce 1986 are synonyms of C. v. variolosus.

Measurements and weights Cacomantis variolosus sepulcralis, Philippines: Wing, M (n ⫽ 24) 110–134 (115.9 ⫾ 2.7), F (n ⫽ 18) 104–120

Brush Cuckoo Cacomantis variolosus 443 (114.7 ⫾ 4.6); tail, M 111–127 (116.8 ⫾ 4.3), F 108–118 (116.2 ⫾ 6.8); bill, M 15.6–18.4 (16.6 ⫾ 1.0), F 14.4–19.1 (17.0 ⫾ 1.3); tarsus, M 15.8–19.2 (17.2 ⫾ 1.0), F 14.8–18.4 (16.7 ⫾ 1.2) (AMNH, CM, CMNH, FMNH, NMP, SMF, ZMUC); C. v. virescens: Wing, M (n ⫽ 12) 101–118 (111.6 ⫾ 5.0), F (n ⫽ 4) 104–111 (108.5); tail, M 120–130 (123.5, n ⫽ 4), F 116–120 (118) (AMNH, BMNH, FMNH); C. v. aeruginosus: Wing, M (n ⫽ 9) 113–121 (116.1 ⫾ 3.4), U (n ⫽ 3) 117–119 (118.0); tail, M 102–125 (119.7 ⫾ 8.24), U 115–130 (123.7) (AMNH, BMNH, MSNG); C. v. infaustus, Moluccas (excluding “heinrichi” as identified by Stresemann (1931) and Ripley (1964)): Wing, M (n ⫽ 24) 115–130 (124.0 ⫾ 4.3), F (n ⫽ 12) 114–130 (122.03 ⫾ 4.0); tail, M 98–124 (109.8 ⫾ 9.4), F 102–123 (114.2 ⫾ 6.7) (AMNH, BMNH, RMNH, USNM);“C. heinrichi” as identified by Stresemann (1931) and Ripley (1964): Wing, M adult (n ⫽ 4) 118–123 (119.7), M juvenile (n ⫽ 3) 108–115 (112.3 ⫾ 10.0); F (n ⫽ 2) 104–115 (108.5); tail, M (n ⫽ 4) 94–114 (104.0), M juvenile 98–112 (100.0), F 85–103 (94.0) (AMNH, YPM); Buru, wing, M (n ⫽ 8) 112–118 (114.8 ⫾ 2.4), F (n ⫽ 3) 113–118 (114.7), tail, M 116–137 (122.7 ⫾ 8.2), F 117–136 (126.3) (MZB); New Guinea: Irian Jaya east to Sepik River: Wing, M (n ⫽ 3) 112– 122 (117.3), F (n ⫽ 2) 117–121 (119), U (n ⫽ 7) 116–122 (119.7 ⫾ 2.1); tail, M 102–108 (105.7), F 100–105 (102.5), U 105–117 (110.1 ⫾ 4.3) (AMNH, ANSP, YPM); D’Entrecasteaux Islands, Dampier islands (Vulcan ⫽ Manam, Rook ⫽ Umboi, Dampier ⫽ Karkar):Wing, M (n ⫽ 16) 114– 130 (122.9 ⫾ 4.8), F (n ⫽ 7) 113–128 (121.±5.2); tail, M 95–118 (106.5 ⫾ 7.2), F 92–102 (100.8 ⫾ 6.9) (AMNH, BMNH, FMNH); C. v. oreophilus, Papua New Guinea, E and S highlands: Wing, M (n ⫽ 23) 110–128 (119.3 ⫾ 4.1), F (n ⫽ 20) 107–120 (116.7 ⫾ 4.7); tail, M 92–114 (108.8 ⫾ 6.9), F 92–102 (95.7 ⫾ 6.7) (AMNH, MCZ, KU,YPM); C. v. blandus, Bismarck Archipelago (Admiralty Islands): Wing, M (n ⫽ 10) 101–117 (112.1 ⫾ 4.4), F (n ⫽ 7) 110–113 (111.4 ⫾ 1.3); tail, M 101–116 (107.5 ⫾ 5.1), F 101–117 (105.9 ⫾ 6.0) (AMNH, USNM);

C. v. macrocercus, Bismarck Archipelago (New Ireland, New Britain, Djaul): Wing, M (n ⫽ 17) 112–139 (121.7 ⫾ 5.8), F (n ⫽ 3) 121–127 (124.3); tail, M 111–124 (126.1 ⫾ 8.1), F 124–129 (126.3); bill, M 15.4–19.1 (17.4 ⫾ 1.1), F 18.0–18.8 (18.3); (Tabar): Wing, M (n ⫽ 2) 126–131 (128.5); tail, M 129–138 (133.5); bill, M 20.3–20.7 (20.5); (Lihir Islands): Wing, M (n ⫽ 11) 117–127 (121.3 ⫾ 3.0); tail, M 110–128 (119.2 ⫾ 4.5); bill, M 16.4–17.1 (16.8) (AMNH, ZMUC); C. v. websteri, Bismarck Archipelago (New Hanover):Wing, M (n ⫽ 8) 115–122 (117.9 ⫾ 2.1), F (n ⫽ 1) 115; tail, M 108–123 (113.6), F 112 (AMNH, ZMUC); C. v. addendus: Wing, M (n ⫽ 12) 115–124 (120.3 ⫾ 3.4); F (n ⫽ 6) 117–126 (120.8 ⫾ 4.1); tail, M 114–138 (121.5 ⫾ 7.0), F 114–125 (119.7 ⫾ 4.8) (AMNH, USNM,YPM); C. v. variolosus, southeastern Australia: Wing, M (n ⫽ 17) 127–148 (135.5 ⫾ 6.1), F (n ⫽ 3) 128–132 (130.3); tail, M 105–128 (115.3 ⫾ 6.1), F 102–113 (106.5) (Higgins 1999); northern Queensland: Wing, M (n ⫽ 12) 122–134 (125.9 ⫾ 3.7), F (n ⫽ 9) 122–127 (124.1 ⫾ 1.7); tail, M 102–112 (105.8 ⫾ 3.6), F 98–107 (102.1 ⫾ 2.7) (AMNH); Northern Territories and Western Australia: Wing, M (n ⫽ 8) 125–131 (128.4 ⫾ 1.9), F (n ⫽ 8) 126–133 (129.5 ⫾ 2.3) (AMNH, RMNH, WAM); Timor: Wing, M (n ⫽ 4) 121–131 (124.5), F (n ⫽ 3) 121–125 (124.0) (Mayr 1944b); Timor: wing, M adult only (n ⫽ 3) 117–125 (123); tail, M 111–117 (114) (ANSP, RMNH, ZMB). Weight, C. v. sepulcralis, Malay Peninsula: U (n ⫽ 10) 28.1–38.0 (Wells 1999); Borneo: M (n ⫽ 1) 30.8 (Thompson 1966); Philippines: M (n ⫽ 21) 24.4–35.5 (32.7), F (n ⫽ 10) 30.2–35 (32.1) (CMNH, FMNH); C. v. infaustus, Buru, M (n ⫽ 4) 27–35 (30) (MZB), Halmahera, M (n ⫽ 2) 33–34 (33.5) (MZB); C. v. oreophilus, New Guinea: M (n ⫽ 9) 29–36 (34.1), F (n ⫽ 5) 35–44 (38.8) (Mayr and Gilliard 1954, Diamond 1972, Diamond and LeCroy 1979), C. v. addendus: M (n ⫽ 8) 34–42 (38.1) (Amadon 1942), M (n ⫽ 9) 38–45.5 (42.2), F (n ⫽ 3) 38–42.5 (40.0) (Galbraith and Galbraith 1962), juvenile F 49.0 (UWBM); C. v. variolosus, eastern Australia south of 22° S:M (n ⫽ 8) 39–49 (41.8); N and W Australia: M (n ⫽ 22) 28–37.5

444 Brush Cuckoo Cacomantis variolosus (33.7), F (n ⫽ 7) 21.8–38.0 (32.1) (Higgins 1999); Western Australia, M (n ⫽ 7) 31–38 (33.8) (WAM). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⱖ 10 ⬎ 1.

Field characters Overall length 24 cm. Adult, a gray-headed cuckoo with a pale throat and underparts and a square tail. Plumage varies from olive brown to light brown above, rufous to buff below, and the eye-ring is yellow in northern part of range and in the Solomon Islands, and yellow to gray in New Guinea, Australia and the Bismarck Archipelago. Females occur in two color phases, one barred and one unbarred. Birds in northern Moluccas vary from gray to brown in plumage. C. variolosus differs from Plaintive Cuckoo C. merulinus within the range shared by both cuckoos in its darker brown upperparts, the head is as dark as the back, the underparts are light brown to rufous, and gray only on throat and upper breast, the rufous underparts grade into grayish cheeks (sharp demarcation of gray and rufous in C. merulinus), and the barring on the tail is incomplete (the pale notches or partial bars do not cross the feather). In Australia and New Guinea, C. variolosus differs from Brush Cuckoo C. flabelliformis in its rusty underparts and short, square tail, and the under tail coverts are rustier than the throat in C. variolosus (throat rustier in C. flabelliformis). Juvenile, barred above and below, crown rufous to tan barred black (not rufous buff with black streaks as in C. merulinus), back darker, throat barred (not streaked), rump barred (not streaked), back with the dark bars wider than the pale bars, and tail with brown to rufous V-shaped notches on edges of feathers with the vane blackish most of length (not buff to rufous bars across feathers). Juvenile is barred below and is not rufous as in juvenile Chestnut-breasted Cuckoo C. castaneiventris.

Voice Two distinct calls: (1) a cadence song or descending song given at a constant speed, a series of melancholy whistled notes, each note rising then quickly dropping at the end “fear-fear-fear . . .”, or

“heet, heet, heet”; the series starts high and becomes lower in pitch or stays on same pitch, about 4 notes in 2 sec, and (2) a shrill rising phrase of three or four notes (“where’s the tea?” or “Piet van Fleet”), the first and third note higher and louder, the phrases stepping up the scale.The songs are similar in Thailand, the Malay Peninsula, Greater Sundas, Lesser Sundas, Sulawesi, New Guinea, New Britain, Timor and Australia (Dahl 1899, Rensch 1931, Stein 1936, Rand 1942a, Rand and Gilliard 1967, Slater 1971, Medway and Wells 1976,White 1984, Beehler et al. 1986, Buckingham and Jackson 1990, Smith 1993a, Thomas and Thomas 1994, Wells 1999, Higgins 1999, Scharringa 1999, Robson 2000a, Supari 2003 NSA, Figure 6.5). In Bougainville and the Solomon Islands, the song is said to differ from song elsewhere (Diamond 1975); no recordings are known. Inskipp et al. (1996) noted an observation of sympatric “sepulcralis” and “variolosus” with different calls in Sulawesi, but all specimens and published observations from Sulawesi are attributed to one form, C. variolosus virescens. In the Moluccas where “heinrichi” occurs, the song begins with a single high whistle, the notes are repeated, increase in loudness and the interval between becomes shorter, and the song ends with a rapid series of whistled notes, the entire song forming a crescendo from a chirp to a high screech (Heinrich 1956), the song apparently “where’s the tea?”.This bird as well as both rufous and gray adults are known as “o pipi todéhe” (on USNM specimen labels), a Tobelo phrase that means “I ask for money”, based on folk lore that this cuckoo’s call portends the arrival of the tax collector, and indicating that locals know the birds as a single species (Taylor 1990). Nevertheless folk taxonomy does not always correspond to biological species, as some other closely related bird species were known by the same name, such as the four species of herons called “o tumara” (Taylor 1990). Song differences may occur between northern and southern cuckoos, with the northern birds (Asian mainland, Greater Sunda Islands, Philippines) having a simpler “fear” call and a shorter “where’s the tea?” call than birds from the central

Brush Cuckoo Cacomantis variolosus 445 Moluccas and New Guinea south through Australia (Figure 6.5), and their songs should be recorded throughout their region. Other calls of C. variolosus in the southern region include a single whistle “fiew” or “fear”, a hiss and an abrupt “churrrt” (Coates and Bishop 1997, Higgins 1999).

Range and status Southeast Asia, resident in lowlands of southern Thailand and peninsular Malaysia, Cambodia and southern Vietnam (Lekagul and Round 1991,Wells 1999, Robson 2000a). Common resident in the Philippines (Kennedy et al. 2000), the Greater Sunda Islands, Lesser Sundas (Lombok, Sumbawa, Flores, Sumba), the Moluccas (Buru, Seram) and Timor (Watling 1983, Coates and Bishop 1997), scarce in Sabah (Sheldon et al. 2001). In New Guinea and northern Melanesia, common resident or local migrant (Gilliard and LeCroy 1961, 1966, Diamond 1972, Coates 1985, 2001, Mayr and Diamond 2001). In Australia, C. v. variolosus are common, migratory in the south, the western birds perhaps resident, others resident to partly migratory to resident, some birds winter in New Guinea (Higgins 1999, Coates 2001). A large, pale adult male taken on Taliabu suggests migration as far north as Sulawesi (SMTD C-45843, Eck 1976, 1977), yet the date (5 October) is late and most birds would have returned to eastern Australia by then; other possible sources of this bird include NW Australia and Timor. Other birds in Taliabu and Sulawesi are small and dark C. v. virescens and C.

v. aeruginosus, and recently-fledged young are small and dark (SMTD C-45842).

Habitat and general habits Broadleafed woodland, in rainforest, secondary forest, logged forest, forest margins, lightly wooded country, scrub, mangroves, thickets of honeymyrtle and paperbark Melaleuca, teak plantations, old rubber plantations, roadsides, cultivation, gardens, often in undergrowth. In the Malay Peninsula, Sumatra, Java and Sabah the cuckoos occur in lowlands, in Kalimantan they are more in hilly country rather than in the lowlands (Holmes 1997). They occur from the lowlands at sea level to 600 m in Malay Peninsula (Wells 1999), to 1700 m in parts of Borneo (Smythies 1999, Sheldon et al. 2001); higher elsewhere, to 2500 m in Sulawesi (Stresemann 1940), 1000 + m in the Moluccas and Lesser Sundas (Rensch 1931, Ottow and Verheijen 1969, Coates and Bishop 1997,Verhoeye and Holmes 1998), 2000 m in the Philippines (Goodman et al. 1995), 1800 m in New Guinea (Coates 2001) and 1200 + m in the Solomon Islands (Diamond 1975). Rufous birds live in montane forest in northern Moluccas, at 1200 m on Gunung Gamkanora, Halmahera, and at 800–1200 m on Gunung Sibela, Bacan (Heinrich 1956, Coates and Bishop 1997). In New Guinea, C. variolosus occur in open wooded country from lowlands to about 1300–2000 m, the upper limit depending on habitat disturbance by man, while Chestnut-breasted Cuckoo C. castaneiventris are in the forest interior in the same altitudinal range and are regular in the upper range but rare at sea level, and Fan-tailed Cuckoo C. flabelliformis are in both second growth and forest interior from 1300 m upward and take over from the other Cacomantis species at higher elevations (Diamond 1972). In the Bismarck Archipelago they are more forest birds than they are in New Guinea (Mayr and Diamond 2001).The cuckoos are abundant in lower and middle tier of open forest in New Britain (Gilliard and LeCroy 1967a), and in secondary forest in the Solomon Islands (Sibley 1951, Blaber 1990, Kratter et al. 2001). In Australia they occur in a wide range of wooded habitats, gallery forests, wet Eucalyptus forests, thickets of paperbark Melaleuca, wattles

446 Brush Cuckoo Cacomantis variolosus Acacia, mangroves, and sometimes even in pines.The cuckoos take prey from leaves and branches within the foliage, and they fly to a lower level and to the ground to catch insects (Higgins 1999,Wells 1999).

Food Insects, mainly caterpillars (hairless and hairy: Pieridae, Lasiocampidae, Saturnidae, Arctiidae), and lepidoptera eggs, grasshoppers, beetles, bugs, wasps, large ants; spiders, snails and fruit (Diamond 1972, Sody 1989, Higgins 1999, KU).

Displays and breeding behavior Males call from regular sites and appear to be territorial, concealed in foliage as they sing. Courtship feeding occurs; male fed a caterpillar to a female (Higgins 1999).

Breeding and life cycle In Java eggs perhaps of this cuckoo species are found in all months from January to November with a peak from April to June (Hoogerwerf 1949), in Flores eggs perhaps of this species are known from July to October (Ottow and Verheijen 1969), in Borneo a fledgling was identified in April (Cranbrook and Wells 1981) and a juvenile was taken in July (SP 17065), and eggs probably this species reported in April (Moyle et al. 2001) and in the Philippines, on Mindanao a female had an egg in late April (BPBM). In Flores they lay from May to October (Verheijen 1964). In Sulawesi a laying female was taken in January (White and Bruce 1986), on Taliabu a short-tailed young cuckoo was taken in October (SMTD C-45842). In the northern Moluccas, cuckoos were in breeding condition in October (Stresemann 1931, Ripley 1964) and a short-tailed chick was taken at Jailolo in late December (USNM 572389). In New Guinea in the Weylands Mt area at Kunupi the cuckoos breed in September (Hartert et al. 1936), in the wet central highlands they breed in the dry season in July and August (Diamond 1972), in the drier Fly River area they breed from November to February (Rand 1941a, AMNH). In dry Australia the cuckoos breed during the rains, in the Kimberley Region of Western Australia from December to March, in N Queensland from Sep-

tember to January and in southeastern Australia from October to February (Pizzey 1980, Storr 1980, 1984b, Higgins 1999). Brood-parasitic. The first record: a cuckoo egg was found in northern Queensland in 1896 in the nest of a Brown-backed Honeyeater Ramsayornis modesta; the nest was “shepherded” until the egg hatched, the young grew its feathers and was nearly ready to fly, and the feathered young was identified as the cuckoo in the Australian Museum (Campbell 1900). Hosts, more than 60 species have been recorded, but not all of these were identified with certainty as this cuckoo’s eggs, and in Java, Flores and Borneo some records may refer to Lesser Sunda Cuckoo Cuculus lepidus, which has eggs similar in size to Cacomantis variolosus sepulcralis. Hosts in Sumatra include tailorbirds Orthotomus species; hosts in Java attributed to this cuckoo include Long-tailed Shrike Lanius schach, Grey-headed Canary-flycatcher Culicicapa ceylonensis, Striated Grassbird Megalurus palustris, Rufous-tailed Fantail Rhipidura phoenicura, Pied Fantail R. javanica, Hill Blue-flycatchers Cyornis banyumas, Thicket Flycatcher Ficedula hyperythra, Sunda Forktail Enicurus velatus and Pied Stonechat Saxicola caprata and others including fruit-eating bulbuls Pycnonotus species (Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967); in Borneo a host is Chestnut-naped Forktail Enicurus ruficapillus (Cranbrook and Wells 1981); in Sulawesi hosts include Olive-backed Sunbird Nectarinia jugularis and Mountain Tailorbird Orthotomus cuculatus (Rozendaal and Dekker 1989, Beisenherz 1998); in Sula Islands host is Black Sunbird N. aspasia (Stones et al. 1997); in Flores hosts are said to be Pied Stonechat and an unidentified dome-nesting warbler or babbler (Ottow and Verheijen 1969); in Buru host is Buru Yellow White-eye Zosterops buruensis (Siebers 1930); in New Guinea hosts are White-shouldered Fairywren Malurus alboscapulatus, Yellow-tinted Honeyeater Lichenostomus flavescens, Lemon-breasted Flycatcher Microeca flavigaster and others (Diamond 1972, Beehler et al. 1986); in New Britain hosts are sunbirds Nectarinia jugularis and N. aspasia (Dahl 1899, Meyer 1933); in the Solomon Islands hosts include a monarch flycatcher Monarcha castaneiventris (Cain and Galbraith 1956, Hadden 1981). In Australia the hosts include fantails Rhipidura, flycatchers

Dusky Long-tailed Cuckoo Cercococcyx mechowi 447 Myiagra, small honeyeaters Ramsayornis and Melithreptus, Purple-crowned Fairy-wrens Malurus coronatus and robins Petroica (both open cup-like nest and closed-nest hosts) (Beruldsen 1980, Brooker and Brooker 1989a). Eggs, in Java three egg morphs have been attributed to this cuckoo. Some eggs are like eggs of shrikes (greenish or grayish with irregular coarse blotches), other eggs like fantails’ (whitish with buffy olive to olive speckles and dots) and the third a “neutral” morph (white with pale brown spots and speckles, like eggs of Sunda Robin Cinclidium diana and Indigo Flycatcher Eumyias indigo), 20 ⫻ 15.3 mm (Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967); in Flores eggs attributed to

the cuckoo are whitish yellow with reddish white spots both large and small, 20 ⫻ 14 mm (Ottow and Verheijen 1969). In Australia, eggs are variable in color and pattern (1) whitish with brownish spots around large end, (2) whitish with few tiny spots of black; each corresponding to a host species; 19 ⫻ 14 mm, weight 1.8 g. Incubation period is 12–13 days. Nestling evicts host eggs and young within two days.The nestling period is 17–19 days, and the cuckoo young is fed by its foster parents for a month after leaving the nest. Nesting success, of 10 cuckoo eggs, 7 hatched and 6 fledged (Campbell 1900, MacGillivray 1914, Beruldsen 1980, Brooker and Brooker 1989a, Strahan 1994, Higgins 1999).

Genus Cercococcyx Cabanis, 1882 Cercococcyx Cabanis, 1882, Journal für Ornithologie , 30, p. 230. Type, by original designation and monotypy, Cercococcyx mechowi Cabanis. Cuckoos with small bodies, large rounded wings and a

very long tail with broad feathers, occurring in forests in Africa. The name refers to the long tail (Gr. kerkos, tail; kokkux, the cuckoo). Three species.

Dusky Long-tailed Cuckoo Cercococcyx mechowi Cabanis, 1882 Cercococcyx mechowi Cabanis, 1882, Journal für Ornithologie, 30, p. 230. (Angola) Monotypic.

Description ADULT: Sexes alike, above, head and back dusky black, wing coverts dusky black with rufous bars, wing brownish black with rufous notches on the outer web of the flight feathers, tail long (more than half the total body length, longer than the wing), graduated, rectrices broad with tips white, T1 with rufous notches but no bars, T2 to T5 edge with rufous notches and vane with interrupted pale bars (nearly continuous in T5) light brown on outer webs and white on inner webs and white spots along the shaft, face with lores and forehead uniformly dark, cheeks dark gray with indistinct white streaks; underparts, throat to breast whitish with black bars, belly and under tail coverts unmarked rufous buff (tail coverts sometimes with an incom-

plete black subterminal bar), tail barred white below, bend of wing white, under wing coverts white, primaries below broadly marked white; eyering yellow to greenish yellow, iris brown, bill greenish black, lower mandible greenish, mouth yellow, feet yellow. JUVENILE: Above, forehead rufous, crown and nape rufous with black base of feathers, back dusky black barred rufous, the bars formed by 2 mm broad tips, wing coverts blackish with broad rufous edges, wing dusky black barred rufous, tail blackish with rufous bars and notches, the rectrices pointed and lack white tips, the face dark with cheeks dark gray and rufous; underparts, throat to upper breast blackish with dark rufous edges, lower breast, belly and under tail coverts white with black spots, under tail black barred rufous, with white on the bars only along the shaft (tail not barred white as in adult) (AMNH 831778).

448 Dusky Long-tailed Cuckoo Cercococcyx mechowi NESTLING: At hatching the skin is black with a yellow rump; the feet are pale (Brosset and Erard 1986). SOURCES: AMNH, BMNH, CM, FMNH, ROM, UMMZ, USNM, ZFMK.

Measurements and weights Wing, M (n ⫽ 18) 127–142 (138.1 ⫾ 4.9), F (n ⫽ 8) 129–142 (134.8 ⫾ 5.6); tail, M 178–211 (190.2 ⫾ 9.0), F 172–194 (183.2 ⫾ 8.0); bill, M 17.8–21.9 (19.5 ⫾ 1.0), F 16.6–22.2 (19.1 ⫾ 1.6); tarsus, M 16.0–18.8 (17.6 ⫾ 0.9), F 17.3–21.0 (18.3 ⫾ 1.3) (AMNH, FMNH, UMMZ). Weight, M (n ⫽ 5) 52.0–57.5 (54.4), F (n ⫽ 3) 52.5–62.0 (57.5) (AMNH, FMNH, UMMZ, ZFMK, Eisentraut 1973, Friedmann 1966, Ripley and Heinrich 1966a). Wing formula, P7 ⱖ 8 ⬎ 6 ⬎ 9 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 31–34 cm. Long-tailed cuckoo with uniform dark gray upperparts, barred rufous belly and under tail coverts, and rufous barring on wings. Distinguished from other long-tailed cuckoos by the dark face and by the darker rufous underparts. The bird illustrated as a juvenile in Irwin (1988) is nearly molted from juvenile to adult plumage; the juvenile plumage in a bird collected recently in the Central African Republic (AMNH), is darker rufous. Distinguished from Olive Long-tailed Cuckoo C. olivinus by voice, and juvenile from young Barred Long-tailed Cuckoo C. montanus by the darker underparts.

Voice Most frequently heard call (1) is a series of three plaintive whistles, “za-so-foé” or “hu-wuk-wuh” or “huit-huit-huit”; the first note soft and slightly downslurred at 1.6 kHz, the second rising to 2 kHz then falling, the third loudest and rising from 1.8 to 2.3 kHz then dropping to 1.7 kHz, each note 0.14–0.18 sec in length and the song 0.8 sec. The long song (2) is a run of 20 to 30 whistled notes lasting 5–10 sec, “teu, tew, du du du..” or “wuhwuhwuhwuh . . .”, the notes in the beginning of series dropping from 2.0 to 1.7 kHz and near the end of

series about 0.4 kHz lower than the earlier notes and the repeat time slowing near the end. Birds also give an excitement call (3), a series of 3 short, strident rising whistles of high pitch (32 to 3.4 kHz). Although Chappuis (1974, 2000) suggested regional differences (species?) between West Africa and Central Africa in song (2), no difference is apparent in the recordings available; and the reported presence of (1) only in West Africa where recorded in southeastern Nigeria and call (3) in Central Africa may be a sampling error, as (1) occurs also in southern Cameroon (Bates 1909), Gabon (Christy and Clarke 2000) and perhaps eastern Zaire (Chapin 1928); call (3) on the other hand has been noted in Gabon to eastern Zaire, Uganda and NW Tanzania (Chapin 1928, 1939, Keith and Gunn 1971, Eisentraut 1973, Chappuis 1974, 2000, Brosset and Erard 1986, L. McBride tape from NW Tanzania). Often calls at night.

Range and status West and Central Africa from Sierra Leone, Liberia, Guinea and Ivory Coast to Ghana, Nigeria and west and southern Cameroon, Gabon, Central African Republic (Dzanga-Sangha Reserve), Congo, E Zaire, Uganda and NW Tanzania, central Zaire and NW Angola (Bannerman 1951, Friedmann 1966, Ripley and Heinrich 1966a, Louette 1981a, Brosset and Erard 1986, Dyer et al. 1986, Grimes 1987, Irwin 1988, Baker and Baker 1994, Christy and Clarke 1994, Elgood et al. 1994, Louette and Herroelen 1994, Cheke and Walsh 1996, AMNH). Resident, population-wide movements are unknown; the birds are silent in nonbreeding season and difficult to see in forest.

Olive Long-tailed Cuckoo Cercococcyx olivinus 449

Habitat and general habits Lowland mature forest, lower levels and dense undergrowth, and montane forest and forest edge, often near water courses (Brosset and Erard 1986); on Mt Nimba they occur to 1200 m (Colston and Curry-Lindahl 1986, Gatter 1987), in Cameroon on Mt Kupé and Rumpi Mts to 1100 m (Eisentraut 1993), and in western and southwestern Uganda and NW Tanzania in lowland forest from 900 to 1600 m, where they call all day and inhabit the midcanopy and undergrowth (Stevenson and Fanshawe 2002).

Food Insects especially caterpillars, most of them hairy, and beetles, ants; spiders, small snails and seeds (Bates 1930, Chapin 1939, Irwin 1988).

Breeding During rains, in lowlands of Liberia they call at night from August to November and February to March or later (Gatter 1997), in Nigeria they call from April to December (Dyer et al. 1986), in Cameroon the gonads are enlarged December to February, in NE Gabon they lay in November and

January (Brosset and Erard 1986), and in Angola they are in breeding condition in October and November (Heinrich 1958). Brood-parasitic. Hosts, a grown, fledged cuckoo of this species (AMNH 159058) was fed by Brown Illadopsis Trichastoma fulvescens (Chapin 1928, 1939); an adult cuckoo perched on a nest of Blue-headed Monarch-flycatcher Trochocercus nitens as it laid (Brosset and Erard 1986); and Forest Robin Stiphrornis erythrothorax are hosts in the Ivindo River area, NE Gabon, with three observations of a brownish cuckoo egg in the nest, two of which hatched cuckoos. One nestling ejected the robin egg the day after the cuckoo hatched; the other nestling ejected the just-hatched young robin, then developed plumage apparently of this cuckoo species (Brosset and Erard 1986). Forest robins of one species or another are common in Dzangha-Sangha Reserve, Central African Republic (Beresford and Cracraft 1999), where a recently-fledged short-tailed cuckoo is known (AMNH), and in the Nindam forest in Nigeria, where the cuckoos occur (Dyer et al. 1986). Eggs are brownish with a darker wreath around the blunt end, the size is not recorded. Incubation period is at least 12 days.The nestling period is unknown.

Olive Long-tailed Cuckoo Cercococcyx olivinus Sassi, 1912 Cercococcyx olivinus, Sassi, 1912, Annalen des K. K. Naturhistorishen Hofsmuseums Wien, 26, p. 31, p. 378. (Rutshuru Plain, Belgian Congo). Monotypic.

primaries below broadly marked white; eye-ring greenish yellow, iris dark brown, bill slate, lower mandible greenish with blackish tip, mouth yellow, feet yellow.

Description

JUVENILE: Upperparts dark brown, head, back and wing covert feathers with rufous tips, feathers paler towards the rump, wing dark brown barred rufous, tail black with rufous notches on outer vane, rectrices pointed and lacking white tips, T5 with buff to whitish bars, underparts, throat and breast dark brown with rufous tips, belly paler (FMNH 296680).

ADULT: Sexes alike, crown dark gray, back dark brown with olive or bronze sheen, unbarred, wing dark brown, tail long, rectrices broad and tipped white or pale rufous, T1 dark brown with trace of buff edges or notched,T2 to T5 with whitish bars, lores to side of forehead speckled with whitish feathers having a dark gray margin, cheeks streaked gray; underparts, throat, breast, upper belly and flanks whitish with blackish bars, lower belly and under tail coverts unbarred whitish buff, tail long, dark with indistinct white and brown bars and spots, bend of wing white, under wing coverts white,

NESTLING: Undescribed. SOURCES: AMNH, BMNH, CM, FMNH, YPM.

450 Olive Long-tailed Cuckoo Cercococcyx olivinus

Measurements and weights Wing, M (n ⫽ 16) 138–157 (145.8 ⫾ 5.5), F (n ⫽ 8) 134–150 (142.8 ⫾ 7.5); tail, M 157–186 (171.4 ⫾ 8.5), F 161–182 (168.6 ⫾ 8.8); bill, M 18.4–20.0 (19.4 ⫾ 0.5), F 19.1–20.1 (19.3 ⫾ 0.9); tarsus, M 16.3–21.0 (18.3 ⫾ 1.1), F 16.3–19.4 (18.4 ⫾ 1.0) (AMNH, FMNH,YPM). Weight, M (n ⫽ 3) 63–68.5 (65.5), F (n ⫽ 1) 56.0 (AMNH, FMNH, YPM, Ripley and Heinrich 1966a). The species was collected long before it was first described. Grauer collected two birds in montane forest in east-central Africa, and sent them to the Vienna Museum. Sassi described the species from these birds, which he compared with specimens in the Berlin Museum. Wing formula, P7 ⬎ 8 ⫽ 6 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 32–34 cm. Long-tailed cuckoo with dark brownish gray upperparts, barred below, with buff under tail coverts and white notches in outer tail feathers. In the hand, bronze-brown above and the barring below less broad and webs of primaries less notched or barred than in Dusky Long-tailed Cuckoo C. mechowi. Juvenile is paler on belly than juvenile C. mechowi.Adult differs from C. mechowi in voice.

Voice Short song, a whistled “ee-eye-ow”, the first note higher and louder, the second note sliding down from the first, somewhat like the “Piet-my-vrow” call of Red-chested Cuckoo Cuculus solitarius but the notes less clearly separated and less emphatic. Long song, a persistent, loud melodious “do-you?”, the long 0.6 sec note sliding slightly down in pitch (the “do-you” as inflected in British English, not in US English!), repeated at rate of 10 phrases in 25 sec, often for long periods with more than 30 phrases (Chapin 1939, Keith and Gunn 1971, Chappuis 1974, 2000, Brosset and Erard 1986, Stjernstedt 1994, Aspinwall and Beel 1998).

Range and status W Africa from Guinea, Ivory Coast, Ghana, southern Nigeria and Cameroon to Gabon, Central African

Republic, northern Congo, western Uganda, northern Zaire, Kasai, NW Angola and NW Zambia (Chapin 1928, 1939, Serle 1954, Louette 1981a, Thiollay 1985, Brosset and Erard 1986, Grimes 1987, Irwin 1988, Gartshore 1989, Louette and Herroelen 1994, Elgood et al. 1994, Rossouw and Sacchi 1998, Dowsett et al. 1999b). Resident. Population density in Gabon, c. 1 in 10 or 15 ha (Brosset and Erard 1986).

Habitat and general habits Forest, mainly unbroken mature forest, and small patches of forest, logged forest, secondary forest, Cryptosepalum forest on Kalahari sands in NW Zambia (Benson and Irwin 1965), and gallery forest. In Liberia they are in northern highlands and southern forests (Gatter 1997), in western Cameroon in lowlands around Kumba (Serle 1954) and from 500–1500 m on Mt Kupé (Bowden 2001), in Gabon throughout primary forest (Brosset and Erard 1986), in Uele in lowlands (Chapin 1939), and in Angola in both primary and secondary forest (Ripley and Heinrich 1966).They occur regionally with Dusky Longtailed Cuckoo C. mechowi through much of the range and are less restricted to large patches of primary forest.They occur together locally with C. mechowi at Tai forest in Ivory Coast, Ivindo forest in Gabon, and Semliki NP, Kibale NP and Budongo Forest Reserve in western Uganda.They are a lowland forest bird in the western rift area of east-central Africa near the montane C. montanus.They perch motionless in midand high strata and crowns of trees, shy and secretive, noticed when they call or when they are in mixed species parties of foraging birds.

Barred Long-tailed Cuckoo Cercococcyx montanus 451

Food Insects, mainly hairy caterpillars (Chapin 1939).

Breeding Apparently during the rains in many regions but between the main rains (little dry season) in wetter regions. In Liberia song is heard in October and November, January to April, and in June, at Mt Wuteve day and night in January (Gatter 1997), in Cameroon at Mt Kupé they call throughout the year (Bowden 2001), in NE Gabon the host nests from December to February (Brosset and Erard 1986, Christy and Clarke 1994), in Uele an egg was found in September (Chapin 1939), and in Angola

the gonads are enlarged from September to November (Ripley and Heinrich 1966a). In Zambia they are heard and observed August to January (Benson and Irwin 1965,Aspinwall and Beel 1998). Brood-parasitic. Host, Pale-breasted Illadopsis Trichastoma rufipennis, based on the probable identification of a cuckoo egg in a nest of this forest songbird in NE Gabon. The egg was deep blue with brown-violet spots around the large end, the egg was larger than the egg of the host and unlike that of other cuckoos in the same area; the egg disappeared (Brosset and Erard 1986).An egg in the oviduct of a laying female at Avakubi, Ituri, was white, unspotted, 23 ⫻ 16 mm (AMNH 159063, Chapin 1928, 1939). Incubation and nestling periods are unknown.

Barred Long-tailed Cuckoo Cercococcyx montanus Chapin, 1928 Cercococcyx montanus, Chapin, 1928, American Museum Novitates, 313, p. 6 (Kalongi, 6900 feet, Butahu Valley, Ruwenzori Range, Belgian Congo) Other common names: Barred Cuckoo. Polytypic. Two subspecies. Cercococcyx montanus montanus Chapin, 1928; Cercococcyx montanus patulus Friedmann, 1928.

JUVENILE: Back feathers edged with rufous, rump slate with rufous feather margins, rectrices pointed, barred with rufous and lacking white tips, throat rufous and blackish streaked white; underparts white with black blotches and streaks (AMNH 450860, ZMUC 59.579, under tail coverts more strongly barred than in adult; iris brown, bill dark brown, feet yellow.

Description ADULT: Sexes alike, crown dark olive brown streaked with buff and rufous, back dull olive brown barred with rufous, the bars formed by 2 mm broad feather tips, wing brown with conspicuous rufous bars on the outer vane and upper coverts, tail long, the rectrices broad, dark brown barred with rufous tan and tipped white or pale rufous, T1 with conspicuous buff and rufous notches,T2 to T5 with conspicuous rufous bars and white spots, lores to sides of forehead are speckled with buff feathers having a dark gray margin, cheeks streaked buff; underparts whitish, throat, breast and belly barred black, under tail coverts buff variably barred with black, tail dark brown barred rufous, bend of wing white, under wing coverts whitish, primaries below broadly marked white; eye-ring yellow, iris brown, bill black, lower mandible greenish at base, feet yellow.

NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, YPM, ZMUC.

Subspecies Cercococcyx montanus montanus Chapin, 1928; smaller, above dark with greenish sheen, face and underparts darker, juvenile throat blackish; montane forests of east-central Africa from Ruwenzori to Lake Tanganyika; Cercococcyx montanus patulus Friedmann, 1928; larger, above paler, face and underparts paler; juvenile throat dusky streaked white; Kenya to southern Zaire, Zambia, E and S Tanzania, Malawi, E Zimbabwe and Mozambique. Specimens at London and Tring and in the Vienna Museum were thought to be juveniles of

452 Barred Long-tailed Cuckoo Cercococcyx montanus C. mechowi, because they were barred on the back, like juveniles of other long-tailed cuckoos. In the Ruwenzori Mts, Chapin heard singing long-tailed cuckoos, and when his collector took two nonsinging birds with somewhat enlarged gonads and plumage that differed from juveniles and adults of the other long-tailed cuckoos, Chapin realized they were a distinct species. He heard the three species each with its own song (Chapin 1928).

Measurements and weights C. m. patulus: Wing, M (n ⫽ 9) 139–148 (144.4 ⫾ 3.4), F (n ⫽ 6) 145–150 (148.0 ⫾ 1.2); tail, M 172–198 (184.4 ⫾ 8.4), F 178–192 (185.6 ⫾ 4.6); bill, M 18.5–20.4 (19.4 ⫾ 0.8), F 18.1–19.9 (18.7); tarsus, M 16.8–21.4 (18.1 ⫾ 1.4), F 15.2–20.0 (17.3 ⫾ 2.2) (AMNH, FMNH,YPM, ZMUC); C. m. montanus: Wing, M (n ⫽ 9) 143–154 (148.8 ⫾ 3.8), F (n ⫽ 8) 143–154 (142.8 ⫾ 4.1); tail, M 175–190 (181.0 ⫾ 5.3), F 167–187 (1 7 7 ⫾ 4.1) (M,F Louette and Herroelen 1994; F AMNH, FMNH). Weight, M (n ⫽ 8) 56.5–60.5 (58.7), F (n ⫽ 3) 52–64 (57.8), U (n ⫽ 1) 56.0 (AMNH, FMNH, YPM, Ripley and Heinrich 1966b, Britton 1977, Blake et al. 1990, Hanmer 1995). Wing formula, P7 ⬎ 6 ⬎ 8 ⬎ 5 ⬎ 9 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 32–34 cm. Long-tailed cuckoo with upperparts brown and barred rufous buff (unbarred in adult Dusky Long-tailed Cuckoo C. mechowi and Olive Long-tailed Cuckoo C. olivinus), wing barred, barred underparts, under tail coverts pale buff with a few brown bars.The long tail and lack of rufous on breast contrast with Red-chested Cuckoo Cuculus solitarius; upperparts brown rather than gray (contrast with Asian and Madagascar Lesser Cuckoos Cuculus poliocephalus and C. rochii). Juveniles are not distinguishable from other long-tailed cuckoo juveniles in the field.

Voice The short song is a 4-note “Piet-Piet-my-vrou” like 3-note song of Red-chested Cuckoo but not as emphatic. The long song “reoo reoo, ree-ree,

ree-ree” starts with notes like the whistled phrase “do-you” of Olive Long-tailed Cuckoo then the song leads into double notes with the first note rising, the second note (louder and longer) sliding down in pitch, with the first note becoming progressively higher from phrase to phrase and the second note progressively lower; the song louder near the end continues as long as 40 sec (Chapin 1928, 1939, Moreau and Moreau 1937, Keith and Gunn 1971, Chappuis 1974, 2000, Stjernstedt 1984, 1993, Jones 1987). Sings at night during moonlight and before sunrise, and during the day when the weather is foggy and humid (Sclater and Moreau 1932, Ripley and Heinrich 1966b, Zimmerman et al. 1996).

Range and status Africa from Albertine montane forests to Lake Tanganyika and East Africa from Kenya south to southern Zaire, Zambia, Zimbabwe, east and southern Tanzania, and Malawi to Mozambique. Resident, with some seasonal migration. In Kivu C. m. montanus have been taken in 10 months and are apparent residents (Schouteden 1968); C. m. patulus with small gonads occur on Idjwi I, Lake Kivu, in July (FMNH). In Kenya they occur on Mt Kenya and in the Aberdare Mts. In Tanzania in Usambara Mt forests they occur mainly from April to September when they sing, and breed there as late as November, and in the Uluguru Mts they are in breeding condition in November and December (YPM); they also occur in as nonbreeding visitors in Kenya in coastal lowlands from September to November (Moreau

Genus Surniculus 453 and Moreau 1937, Britton 1977, Evans 1997, Evans et al. 1997, Lewis and Pomeroy 1989, Louette and Herroelen 1994, Zimmerman et al. 1996, Stevenson and Fanshawe 2002). In southern Tanzania they occur and apparently breed from October to January (Stjernstedt 1970, ZMUC), in Zambia recorded in Gwembe Valley from November to February (Aspinwall and Beel 1998, Dowsett et al. 1999a), in southern Africa from October to May (Benson and Benson 1977,Vernon et al. 1997). Uncommon, secretive and overlooked except when singing. Records in southern Africa are mainly in Zimbabwe at the Haroni-Rusitu river confluence before the forest was cleared (Vernon et al. 1990, 1997) and in the lower Zambezi valley ( Jones 1987, Irwin 1987a), also in Mana Pools NP (Hustler 1985).

Habitat and general habits Forest, montane in most areas, in East Africa also in riparian and lowland forest, coastal evergreen forest and thickets and forest-savanna mosaic, and miombo (Chapin 1928, 1939, Benson and Benson 1977, Rowan 1983, Irwin 1988). In Burundi and Uganda they occur at 1950 and 2075 m (FMNH), they are locally common between 1700 and 2100 m on E slopes of Mt Kenya and S slopes of the Aberdare Mts, and between 900 and 1600 m in the Usambara Mts (Zimmerman et al. 1996).They call from the canopy of tall trees, often at night; they feed at all levels of the forest (Stevenson and Fanshawe 2002).

Food Insects, mainly hairy caterpillars, taken in foliage of trees; occasionally small snails (Irwin 1988).

Breeding Breeds during rains, there are few dates: on Mt Kenya a female had an oviduct egg in March (Keith 1968), in Burundi a bird had an oviduct egg in August (FMNH 357946), in the Usambara Mts birds are in breeding condition from September to December and at Amani a female was laying in November (Moreau and Moreau 1939), and in the Uluguru Mts a short-tailed juvenile (ZMUC 59.579) taken in late February indicates breeding in January; in Malawi they call from December to April (Benson and Benson 1977), and in southern Africa they call at night in January (Chittenden 1995). Brood-parasitic. Hosts, suspected to be Sharp’s Akalat Sheppardia sharpei and other akalats, and African Broadbill Smithornis capensis. A female cuckoo dismantled an akalat’s nest and removed or destroyed the akalat’s eggs (Moreau and Moreau 1939). An egg laid by a cuckoo that flew into a room and was held overnight was white with a faint reddish band around the thick end, similar to the band of blotches on an akalat egg but paler, “a ghost of the egg of Sheppardia” (Moreau and Moreau 1939), and an oviduct egg was unmarked white like broadbill egg, 21 ⫻ 15 mm (Dean et al. 1974). A broadbill nest held a white egg with faint band of brownish spots around the thick end (23 ⫻ 17 mm) along with a normal white broadbill egg, and later the nest held a young cuckoo, identified by comparison with the egg (Moreau and Moreau 1939). An egg in the oviduct of a female (FMNH 357946) in Kibira NP, Burundi, was 24 ⫻ 13 mm. Incubation and nestling periods are unknown.

Genus Surniculus Lesson, 1830 Surniculus Lesson, 1830, Traité d’Ornithologie, livre 2, p. 151. Type, by subsequent designation Cuculus lugubris Horsfield. The name is a diminutive of French le Surnicou, a name given to drongo cuckoos by Lesson (genus Surnia, owl + genus Cuculus,

cuckoo), a term that corresponds to Horsfield’s Latin lugubris for mournful, after the somber black plumage. Black plumage, round nostril, the tail forked to square, the outer tail feather T5 short. Four species.

454 Fork-tailed Drongo-cuckoo Surniculus dicruroides

Fork-tailed Drongo-cuckoo Surniculus dicruroides (Hodgson, 1839) Pseudornis Dicruroïdes Hodgson, 1839, Journal of the Asiatic Society of Bengal, 8, p. 136 and plate. (mountains of Nepal) Polytypic. Two subspecies. Surniculus dicruroides dicruroides (Hodgson 1839), Surniculus dicruroides barussarum Oberholser 1912. Other common names: Drongo Cuckoo (part), Asian Drongo Cuckoo (part), Fork-tailed Cuckoo, Indian Drongo-cuckoo.

Description ADULT: Sexes alike, plumage black with steel-blue gloss, head and upper back less glossy, nape often with small concealed white patch at base of feathers, tail feathers tipped black, the two inner pairs of rectrices T1, T2 shorter than the next two rectrices, T3 with outer web broad at tip,T4 curved, the outer vane narrow near the tip, the feather narrow and bent outward giving the tail a forked appearance, the outer tail feather T5 short with narrow oblique white bars; underparts black, leggings white (long feathers on leg, outer vane of feather white, inner vane black or white), under tail coverts with thin whitish bars, bend of wing black, under wing coverts black, white bar across wing base inner webs of P1–7 and S1–4 as seen from below, P8 unmarked black, P9 with white spot near the base of inner web; skin around eye blackish, iris brown, bill black, feet dark gray. JUVENILE: Dull black with white spots on head, wing and breast, white patch on nape, tail feathers tipped white, tail forked less deeply than in adult, bend of wing with white barred feathers. NESTLING: Skin color not described, mouth lining orange (Hume 1888). SOURCES: AMNH, ANSP, BMNH, BPBM, CUM, FMNH, MCZ, MZB, NSM, RMNH, ROM, SMF,TISTR, UMMZ, USNM,YPM, ZRC.

Species and subspecies of drongo-cuckoos Surniculus dicruroides dicruroides (Hodgson 1839); wing and tail long; southern Asia from Himalayas and Yao Shan through Indian subcontinent;

Surniculus dicruroides barussarum Oberholser 1912; wing and tail shorter; Assam, northern Burma, northern Thailand, Laos, Cambodia,Vietnam,Yunnan and Hainan. Drongo-cuckoos in Asia were known in the early 19th century as two species, Fork-tailed Drongocuckoo S. dicruroides and Square-tailed Drongocuckoo S. lugubris. Blyth (1845) pointed out that some northern birds had a short unforked tail, and some southern birds had a long forked tail, and blithely assumed they were all one species. Hume found no consistent differences between northern and southern birds, and he concluded that they were conspecific (Hume and Davison 1878). He noted that some birds have a more forked tail than others, and some have longer and stouter bills and others slender and shorter bills, and that these characters do not co-vary with sex. Hume’s approach was to compare birds between different geographic regions, and not to entertain the idea that there may be two species living together in one region. In fact, when comparisons are restricted to the breeding season, and the migrant and wintering birds are considered in their own season, and when songs are available, the variation in museum specimens is consistent with the idea of two species of drongo-cuckoos breeding on the Asian mainland, and of more than one taxon wintering together. Birds in northeast India and in Sri Lanka during the breeding season indicate two kinds together without intermediate specimens.Also, the songs available on field recordings indicate two kinds of drongocuckoos, Fork-tailed Drongo-cuckoo and Squaretailed Drongo-cuckoo in southern and southeast Asia, and apparent sympatry in at least one region (northeast India). The fork in the tail is formed by the outer feather T4 being curved and longer than the inner feathers; the depth of the fork is measured on the straightened feather from the tip of T1 to tip of T4, in adult tails (no white tip on the rectrices). Drongo-cuckoos in India and Southeast Asia sort reasonably well into those with forked tails and those with square tails, even within the breeding season. In their breeding areas, in S. d. dicruroides the outer T4 feathers are

Fork-tailed Drongo-cuckoo Surniculus dicruroides 455 10 mm or longer than T1, in S. d. barussarum they are 412 mm and in S. lugubris the outer feathers are 0 to 6 mm longer. In S. dicruroides the outer feather T4 is often narrow at the tip with hardly any outer vane, and it curves outwards at the tip; in S. lugubris T4 is broad at the tip and usually straight. Fork-tailed Drongo cuckoos are generally more glossy; the two drongo-cuckoos are otherwise similar in color and morphology. Hodgson’s (1839) name dicruroides of the northern bird describes the drongo Dicrurus-like appearance of the cuckoo. One type specimen is at Tring, (BMNH 43.1.13.1008) (Hodgson no. 502), wing 141, tail 143, tail fork 15; others are at CUM (as in Benson 1999), 19/Cuc/34/a/5 (Strickland no. 2124), wing 131, tail 147, tail fork 12.2; and 19/ Cuc/34/a/7 ( Jardine no. 2282), wing 136, tail 144, tail fork not measurable as T4,T5 are missing. The measurements are within the size range of the type description (Hodgson 1839). This similar plumages of drongo-cuckoos and drongos led to the idea that cuckoos were mimics of the drongos (Baker 1906), and a few field observations suggested they parasitized drongos (Hume and Davison 1878, Baker 1906, Makatsch 1955). However, there is no compelling evidence. For example, the geographic variation of the cuckoos’ plumage does not parallel the variation of drongo plumage.The drongo with the most similar plumage is Black Drongo Dicrurus macrocercus which lives with drongo-cuckoos in Himalayan Asia, but this drongo also lives in the Greater Sunda Islands where other species of drongo-cuckoos have shorter, square or less forked tails. No species of drongo has populations with a long forked tail in the north and a shorter more rounded tail in the south (Mayr and Vaurie 1948, Vaurie 1949). Geographic variation in the drongo-cuckoos is independent from that within or among the drongo species. The type series of Surniculus dicruroides barussarum Oberholser 1912 taken in February on Tana Bala I, west of Sumatra, and the type series of S. lugubris massorhinus Oberholser 1924 taken in January in the Siak River area near the Straits of Malacca, Sumatra, both in USNM, are wintering S. dicruroides, and are similar in appearance to breeding birds of Assam,Yunnan and southeast Asia.

Measurements and weights S. d. dicruroides, Nepal, Sikkim and northern India (Himachal Pradesh to Madhya Pradesh): Wing, M (n ⫽ 23) 125–146 (137.2 3 ⫾ 6.1), F (n ⫽ 7) 128–145 (135.9 ⫾ 6.4); tail, M 124–147 (133.9 ⫾ 7.4), F 122–143 (130.3 ⫾ 8.3); bill, M 18.3–22.0 (19.8 ⫾ 1.0), F 15.7–20.8 (18.7 ⫾ 1.4); tarsus, M 14–16.5 (14.6 ⫾ 0.9), F 13.9–15.9 (15.0 ⫾ 0.6) (AMNH, FMNH, UMMZ); China: Wing, M (n ⫽ 6) 138–143 (140.3 ⫾ 1.6), F (n ⫽ 1) 138; tail, M 130–142 (130.0 ⫾ 5.1), F 132 (AMNH, MCZ); S. d. barussarum, Assam (Lushai Hills, Garo Hills, Khasi Hills, Cachar), wing, M (n ⫽ 15) 132–144 (137.5 ⫾ 3.5), F (n ⫽ 4) 129–136 (132.4); tail, M 126–140 (132.5 ⫾ 6.1), F 115–128 (118.8) (FMNH, UMMZ); Laos and Vietnam, wing, M (n ⫽ 14) 133–140 (137.1 ⫾ 2.9), F (n ⫽ 7) (134.0 ⫾ 3.3); tail, M 130–140 (135.8 ⫾ 3.7), F126–140 (133.5 ⫾ 6.0) (BMNH, FMNH, UMMZ, USNM,YPM). Weight, F (n ⫽ 1) 50.5 (USNM), longer-winged autumn migrants in Malay Peninsula, U (n ⫽ 41) 32.5–43.3 (Wells 1999). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⱖ 1 0 ⬎ 2 ⬎ 1; in other birds, . . . 2 ⬎ 1 ⬎ 10.

Field characters Overall length 25 cm. Black bird with white pants, the pants formed by long tufted feathers on the posterior surface of the upper legs, under tail coverts barred. A slender black cuckoo with tail broad at tip, long and forked; it differs from Squaretailed Drongo-cuckoo S. lugubris in the longer and more forked tail, and in song in the breeding season. It differs from drongos in its white leggings, white on the under tail coverts and inconspicuous white bars on the tail, the thin decurved bill, black iris (red in drongos), and a white bar on underside of the wing in flight.

Voice Song, five or six loud clear piping whistles mostly rising up the scale, “one, two, three, four, five, six”; the second note lower or at least not higher than the first. Songs were said to be the same in Burma and Borneo (Osmaston 1916, Smythies 1940, 1949, 1960, 1981) and elsewhere through the range of

456 Fork-tailed Drongo-cuckoo Surniculus dicruroides drongo-cuckoos from India and Sri Lanka to southeast Asia, the Greater Sunda Islands and Palawan (Legge 1880, Ali and Whistler 1937, Betts 1947, Ali 1953, King and Dickinson 1975,White 1984, Lekagul and Round 1991, MacKinnon and Phillipps 1993,Wells 1999). However, songs of northern birds differ from those of southern birds, the Square-tailed Drongo-cuckoo S. lugubris. Songs in India through northern southeast Asia including Tonkin and southern China to Hainan have the second note beginning lower or at the same pitch as the first note, and the notes are piping (not mellow) (Figure 6.3); both features are easily heard in the field. Drongo cuckoos north of the Brahmaputra River at Nameri NP in Assam, and in W Arunachal Pradesh also give this song (RBP).The cuckoo also has a long bubbling “crescendo”, a shrill call that quickens and rises then falls away, a variable series of burry or buzzy rising notes ending with descending notes of the same harsh tonal quality given in excitement contexts much as in other drongocuckoos.The crescendo call varies within a bird, in contrast to whistled song which is repeated with little variation. When making the crescendo, the bird spreads and lifts both wings above the back (Ali and Ripley 1969, NSA 15023). It gives the call when excited, as when is pursued by another bird (Hume 1888). Drongo-cuckoos also give a plaintive “wheewheep”—the second note higher than the first— like a drongo (Baker 1906, Smythies 1940), and it is uncertain who mimics whom, as drongos often mimic the calls of other birds (Ali 1949, 1996).

Range and status India from the Western Ghats (Coonoor, Siddapur), Maharastra, Andhra Pradesh (Eastern Ghats at Bhadrachallam) and Madhya Pradesh to the Himalayas, and in Nepal, Bhutan, Burma, Laos, northern Vietnam, southeastern China (Yunnan, Tenggueh) and Hainan (Guangdong et al. 1983, Cheng 1991, Deignan 1945, 1963, Smythies 1960, 1986, Robson 2000a, AMNH). In Sri Lanka, some birds taken in May, July, August and September (BMNH, MCZ, YPM) appear to be this species; other fork-tailed birds in Sri Lanka were taken in winter and may be migrants from the mainland. Seasonally migratory and nomadic in the north

from Himalayas and southern China, and are resident in peninsular India, southeast Asia and perhaps Sri Lanka. Resident from the Western Ghats, Madhya Pradesh, Maharashtra and Andhra Pradesh in the Eastern Ghats, and locally in Assam and eastern Bangladesh, perhaps Sri Lanka, and in Burma (Chin Hills), Laos, Vietnam, Cambodia, southern Yunnan and Hainan. Outside the northern hill region and Assam, only 14 specimens of adult drongo cuckoos taken in India during the breeding season have information on sex. S. d. dicruroides are in Uttar Pradesh (Garhwal), in Nepal (Hetora), in the hills of northern Mahda Pradesh east to Raipur, in Andhra Pradesh (Gadavery Valley), in Mysore in the Western Ghats (Kanara) and in Assam (Cachar). In Burma they sing in the Chin Hills; winter records in Burma are in the north (Katha district), the Toungoo and Pegu Yomas, and Tenasserim. Migrant northern birds occur within the range of southern Square-tailed Drongo-cuckoos S. lugubris in migration and winter, when both species are in the Malay Archipelago as far south as Java. S. dicruroides occur with the monsoon rains in northern India from northern Uttar Pradesh and Himachal Pradesh eastward to Nepal, Sikkim, Bhutan and Arunachal Pradesh, and in southern China (Sichuan, Yunnan, Fujian) (David and Oustalet 1877, La Touche 1931-34, Cheng 1991, Grimmett et al. 1999). In all seasons, large forktailed birds appear in Sri Lanka, and in migration and winter they occur in the Malay Peninsula, Cambodia, Nias and Sumatra, Natuna Is, and Java

Fork-tailed Drongo-cuckoo Surniculus dicruroides 457 (but not Borneo) (AMNH, ANSP, BMNH, MZB, RMNH, YPM); and there is a record in Japan (Brazil 1981). Long-distance migration is indicated by birds attracted to lights in autumn passage over the Main Range at Fraser’s Hill in the Malay Peninsula (the earliest records are 25 September), birds flying at night into a lighthouse in the Malacca Straits (Robinson and Kloss 1922, Wells 1999), and by the appearance of large fork-tailed birds in winter in the range of square-tailed S. lugubris. Smaller fork-tailed birds apparently S. d. barussarum occur in winter in Sumatra and less commonly in Borneo and Palawan. Traced by song, Fork-tailed Drongo-cuckoos occur as breeding birds in northern India (Himachal Pradesh NP, P. Singh recordings; W Assam at 100 m in Nameri NP, RBP; NE Assam, Dig Boi, NSA 100506; W Arunachal Pradesh near Barakpong at 300 m, Sessa at 1220 m, Seppa Road at 1560 m an 17 km N Bomdila at 1649 m, RBP), Nepal (Kathmandu Valley, NSA 12449,White 1984; Chitwan NP, NSA 40913), Bhutan (Connop 1995), western Burma (Mt Victoria, Chin Hills, P. Alström recordings), northern Vietnam (C. Robson recordings), southern Yunnan at Menglun (P. Round recordings) and Hainan (P. Alström recordings). S. dicruroides and S. lugubris occur together in northeastern India (Assam, Arunachal Pradesh), as indicated by song recordings and museum specimens. Field observations and tape recordings are needed to determine the geographic ranges and the extent of sympatry of drongo-cuckoos in southern and southeast Asia.

Habitat and general habits Mainly hill forest, also on the plains. They occur in forests and scrub, foliage canopy, open forests and bamboo jungle (Inglis et al. 1920, Ali and Ripley 1969, Grimmett et al. 1999). Their altitudinal range in India is from 200 to 1800 m, in Yunnan from lowlands to 2100 m (Rothschild 1926) and in Burma

from the lowlands to 1200 m (Smythies 1986).Arboreal, they perch on charred stumps and saplings in burned clearings. Their flight is smooth and direct, not turning and twisting, swooping and acrobatic like that of a drongo; the birds feed on small branches and leaves in the canopy and they also hawk insects like a drongo (Ripley 1950, Ali 1962).

Food Mainly insects, especially caterpillars, soft-bodied insects; occasionally fruit, including figs (Ficus spp.) (Ali and Ripley 1969).

Breeding In Nepal they lay in June, in India they sing and breed in the north during the monsoon from March to October (females in Sylhet had oviduct eggs in April and June), in Sikkim they breed from April to June, and in northern Burma near Myitkyina they sing in April (Stanford and Mayr 1941, Biswas 1960, Ali 1962, 1996, Ali and Ripley 1969, Inskipp and Inskipp 1985, AMNH, BMNH). In Sikkim, adults are in molt in August and September (BMNH). Brood-parasitic. Hosts are mainly babblers; in India the drongo-cuckoos use Nepal Fulvetta Alcippe nipalensis and Dark-fronted Babbler Rhopocichla atriceps, and forktails Enicurus, bulbuls Pycnonotus aurigaster, shrikes Lanius species and Striated Warbler Megalurus palustris (Baker 1906, 1934, Ali and Ripley 1969, Becking 1981). Eggs in nests of babbler hosts and identified by Baker (1934) as Banded Bay Cuckoo Cacomantis sonneratii were re-identified by Becking (1981) as eggs of a drongo-cuckoo.The egg is white with fine purple splotches (oviduct egg of a cuckoo in Nepal) (Biswas 1960, FMNH 229749). Eggs in India are 23.5 ⫻ 17.5 (Baker 1934). The nestling cuckoo evicts the eggs and nestlings of the host (Becking 1981). The incubation and nestling periods are unknown.

458 Philippine Drongo-cuckoo Surniculus velutinus

Philippine Drongo-cuckoo Surniculus velutinus Sharpe, 1877 Surniculus velutinus Sharpe, 1877, Transactions of the Linnean Society of Zoology, (2), 1, p. 321. (Malamaui, Philippines). Polytypic. Three subspecies. Surniculus velutinus velutinus Sharpe, 1877; Surniculus velutinus suluensis Salomonsen 1953; Surniculus velutinus chalybaeus Salomonsen 1953.

Description ADULT: Sexes alike, head, upper back and breast velvety (not glossy) purplish black, nape with small concealed white patch, wing black with purplish, blue or green gloss, tail black with purplish, blue or green gloss, tail feathers T3, T4 with outer web broad at tip tending to bend the feather outward, tail square or slightly emarginate (not forked), outer tail feathers T5 shorter than the others and with narrow oblique white bars; underparts black, leggings white (long feathers on leg, outer vane of feather white, inner vane black or white), under tail coverts black (not barred), under wing coverts black, white bar across wing base inner webs of P1–7 and S1–4 as seen from below, P8 unmarked black, P9 with white spot near the base of inner web; skin around eye dark bluish gray, iris brown, bill black, feet dark gray. JUVENILE: Plumage, either unmarked rusty brown, or black with white spots, varying geographically. NESTLING: Undescribed. SOURCES: AMNH, BMNH, BPBM, CM, CMNH, DMNH, FMNH, JFBM, KU, MCZ, MVZ, NMW, RMNH, ROM, SMF, UMMZ, USNM, UWBM,YPM, ZMUC, ZRC, ZSM.

Subspecies and systematic comments Philippine Drongo-cuckoos were often considered a subspecies of S. lugubris, although their song is different, as is the juvenile plumage (in part) as noted by McGregor (1909). S. velutinus and S. lugubris were regarded as distinct species, in McGregor (1909), Baker (1919) and Riley (1924); they were

later combined as a single species in Stresemann (1940) and Peters (1940). Surniculus velutinus velutinus Sharpe, 1877; adult, head, upper back, throat and upper breast dull velvet black, back and wing glossy steel-blue; juvenile, unmarked rusty brown; the southern Philippines (Basilan, Biliran, Leyte, Samar, Mindanao, Malamaui, Bohol); Surniculus velutinus chalybaeus Salomonsen, 1953; adult, head, upper back and underparts glossy black, southern Philippines (Luzon, Mindoro, Negros); Surniculus velutinus suluensis Salomonsen, 1953; adult plumage velvety black as in S. v. velutinus, wing larger juvenile, black with white spots; Sulu Archipelago (Sulu Islands,Tawitawi, Bongao); On Mindoro, S. v. mindoroensis Ripley and Rabor 1958 was said to have a longer wing and tail than birds on Luzon, and more gloss on the head, neck, upper back and underparts. However, birds on Mindoro do not differ consistently in size from birds on Luzon, when more specimens are compared. The plumage gloss does not appear to be consistently distinct from birds on Luzon and Negros, and the taxon is a synonym of S. v. chalybaeus Salomonsen 1953. Although most juvenile drongo-cuckoos in the Philippine Is (excluding Palawan) are unspotted rufous brown, a few have the white-spotted black juvenile plumage as in mainland S. lugubris.The four specimens in part or full juvenile plumage from Luzon, Mindoro and Negros are black with white spots; the 14 from Leyte, Bohol, Mindanao and Basilan are rufous brown, and the three from the Sulu Is are brown with buff tips on the feathers.There is no evidence in the juveniles or in the adult plumage for the occurrence of two species of drongo-cuckoos within a local region of the Philippines.

Measurements and weights S. v. velutinus, Mindanao and Basilan: Wing, M (n ⫽ 36) 111–130 (117.2 ⫾ 4.0), F (n ⫽ 25) 115– 121 (117.1 ⫾ 2.5); tail, M 86–118 (101.6 ⫾ 6.0), F 92–118 (105.0 ⫾ 7.0); bill, M 17.0–21.9 (19.1 ⫾ 1.3), F 17.0–20.6 (18.7 ⫾ 1.2); tarsus, M 14.0–18.0

Philippine Drongo-cuckoo Surniculus velutinus 459 (15.5 ⫾ 1.0), F 13–18.5 (16.1 ⫾ 1.2) (AMNH, BMNH, BPBM, CM, DMNH, FMNH, JFBM, RMNH, SMF, UMMZ, USNM,YPM); Leyte and Biliran: Wing, M (n ⫽ 6) 114–118 (116.5), F (n ⫽ 4) 110–118 (115.0); tail, M 84–98 (94.2), F 90–108 (96.8) (AMNH, CMNH, DMNH); S. v. chalybaeus, Luzon: Wing, M (n ⫽ 15) 121– 131 (126.8 ⫾ 2.8), F (n ⫽ 6) 124–131 (126.8 ⫾ 2.4); tail, M 98–121 (109.9 ⫾ 5.9), F 92–116 (109.2 ⫾ 7.8) (AMNH, DMNH, FMNH, UMMZ, UWBM, YPM); Mindoro: Wing, M (n ⫽ 5) 128–136 (132.1 ⫾ 2.7), F (n ⫽ 3) 124–130 (128.0); tail, M 104–133 (124.0 ⫾ 11.5), F 112–128 (120) (AMNH, FMNH, MCZ, SMF,YPM); S. v. suluensis, Sulu Islands (Sulu, Tawitawi, Bongao): Wing, M (n ⫽ 3) 125–130 (128.5), F (n ⫽ 6) 120–129 (123.8 ⫾ 3.0); tail, M 105–126 (113.0), F 111–128 (120.3 ⫾ 6.4) (CM, DMNH, ZMUC); Weight, M (n ⫽ 12) 31.4–42 (33.2), F (n ⫽ 7) 36–47 (38.2) (BPBM, CMNH, FMNH, USNM, UWBM,YPM). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⱖ 10.

Field characters Overall length 24 cm. Black bird with white pants. A black arboreal cuckoo with contrasting head and underparts dull velvet and upperparts glossy, tail square, white leggings and black under tail coverts. The velvety body plumage lacks the metallic green color of Square-tailed Drongo-cuckoo S. lugubris and Fork-tailed Drongo-cuckoo S. dicruroides; wings and tail are glossy blue or green. In flight, white bar on underside of wing. Looks like a drongo, but the bars in tail and the fine bill distinguish it as a cuckoo. In south of the range the juvenile is uniformly rufous, unlike other small cuckoos.

Voice Song a series of 8–10 whistles rising in pitch through the series, the song longer, with more notes and less rise in pitch than in Square-tailed Drongo-cuckoo S. lugubris on Palawan (Figure 6.3). Songs of S. velutinus in the Philippines (Luzon and Mindanao) are longer with less rise in pitch than in Asian drongo-cuckoos S. dicruroides and S. lugubris, while

songs of birds in Sulawesi and Halmahera S. musschenbroeki are similar with each note in the series higher, but the individual notes drop in pitch (Figure 6.3). In addition to the whistled song, Philippine Drongo-cuckoos have a burry crescendo call similar to that of other drongo-cuckoos.

Range and status Philippines from Luzon to Mindanao and the sulu Archipelago. Resident. (Gilliard 1950, Salomonsen 1953, Rand and Rabor 1960, Dickinson et al. 1991).

Habitat and general habits Primary dipterocarp forest, secondary forest, wooded country, mixed bamboo woodland; found from sea level to 1000 m.The cuckoos live in forest canopy. Because forests have been nearly eliminated throughout the range of the Philippines where these cuckoos occur, the continued existence of the species is threatened, although the species is not listed in recent evaluations (Collar et al. 1994, BirdLife International 2001).

Food Insects (UWBM).

Breeding Females were taken with oviduct eggs on Mindoro in May and on Bohol in April and May (Ripley and Rabor 1958, Rand and Rabor 1960, Dickinson et al. 1991). Presumably brood-parasitic, the hosts are unknown. The eggs have not been described, and the incubation and nestling periods are unknown.

460 Square-tailed Drongo-cuckoo Surniculus lugubris

Square-tailed Drongo-cuckoo Surniculus lugubris (Horsfield, 1821) Cuculus lugubris Horsfield, 1821, Transactions of the Linnean Society of London, 13(1), p. 179. ( Java) Polytypic. Three subspecies. Surniculus lugubris lugubris (Horsfield 1821), Surniculus lugubris brachyurus Stresemann 1913, Surniculus lugubris stewarti Stuart Baker 1919. Other common names: Drongo Cuckoo (part), Asian Drongo Cuckoo (part).

Description ADULT: Sexes similar (female slightly duller), upperparts black with blue and green gloss, nape often with small concealed white patch, tail square or slightly forked, the feathers tipped black, the two inner pairs of rectrices T1,T2 equal in length to or slightly shorter than the next two rectrices, T3 and T4 with outer web broad at tip, the feather straight and not bent outwards, the outermost tail feather T5 short with narrow oblique white bars; underparts black, leggings white (long feathers on leg, outer vane of feather white, inner vane black or white), under tail coverts with narrow whitish bars, bend of wing black, under wing coverts black, white bar across wing base inner webs of P1–7 and S1–4 as seen from below, P8 unmarked black, P9 with white spot near the base of inner web; skin around eye blackish, iris brown to dark gray, bill black, feet dark gray. JUVENILE: Dull black with white spots on head, wing and breast, white patch on nape, tip of upper tail coverts white, tail square, the feathers rounded with white on tip, sometimes with white spots along shaft, bend of wing with white barred feathers. NESTLING: Skin color not described, mouth lining red, feet vinaceous. SOURCES: AMNH, ANSP, BMNH, BPBM, CM, CU, CUM, DMNH, FMNH, MCZ, MNHN, MSNG, MZB, NMW, NSM, RMNH, ROM, SMF, SMTD, TISTR, UMMZ, USNM, WFVZ, YPM, ZMA, ZMB, ZMUC, ZRC, ZSM.

Subspecies Surniculus lugubris lugubris (Horsfield, 1821); plumage dull, tail short, no fork or fork less than 4 mm; Java, Bali; Surniculus lugubris brachyurus Stresemann 1913; plumage dull, tail shorter than northern birds, no fork or fork less than 5 mm; Malay Peninsula, Borneo, Sumatra and Palawan; Surniculus lugubris stewarti Stuart Baker 1919, plumage similar to S. l. lugubris, bill short; Sri Lanka. In Java, the type locality of S. lugubris, nearly half of 102 specimens lack data on month of collection and nearly half lack information on the sex. Nearly all drongo cuckoos in Java are small (wing ⬍ 130 mm, n ⫽ 92), dull (4 are glossy) and have a nearly square tail; the wing and tail are nearly equal in length. Stresemann (1913b) described S. l. brachyurus from the Malay Peninsula, differentiating it from S. lugubris of Java, Bali and Sri Lanka by relative length of wing and tail, the wing longer in birds from Java, Bali and Sri Lanka, and shorter in birds from the Malay Peninsula, Sumatra and Borneo and nearby islands. Stresemann suggested that birds in Burma were intermediate between brachyurus and dicruroides; but he did not distinguish birds by date, and some may have been passage or wintering migrants from the north. The holotype of S. l. brachyurus, AMNH 6248702, taken in June, has wing 125, tail 112. Kloss (1918) thought brachyurus was the resident form in the Malay Peninsula, as did Wells (1999), and Robinson and Kloss (1923b) recognized brachyurus as the form in Sumatra. Birds in Sumatra taken during the months when northern migrants are absent, are similar to the drongocuckoos in the Malay Peninsula. Birds in Palawan and Balabac, described as S. l. minimus Stuart Baker 1919, do not differ consistently in size from S. lugubris in the Malay Peninsula and Borneo.The holotype, taken in June, has the wing 124, the tail 115 (BMNH 1911.11.16.127). Birds in Sri Lanka described as S. l. stewarti Stuart Baker 1919 are similar in plumage to birds in Java, as noted by Stresemann (1913b); however as

Square-tailed Drongo-cuckoo Surniculus lugubris 461 noted in a small sample by Ripley (1946) the bill is shorter in Sri Lanka than in Java. A larger series shows overlap with birds in Java. The holotype, taken in August, has the wing 125, the tail 146 (BMNH 98.12.2.297).

Measurements and weights S. l. lugubris, Java, April–September: Wing, M (n ⫽ 17) 125–133 (127.4 ⫾ 2.3), F (n ⫽ 5) 127–129 (128.0 ⫾ 1.0); tail, M 120–136 (128.8 ⫾ 4.6), F 118–128 (123.2 ⫾ 4.6); bill, M 19.0–22.0 (20.7⫾ 0.9), F 19.0–21.5 (20.7 ⫾ 1.1); tarsus, M 14.1–17.6 (15.7 ⫾ 1.1), F 14.4–16 (15.7 ⫾ 1.4) (AMNH, MCZ, RMNH); S. l. stewarti, Sri Lanka: Wing, M (n ⫽ 18) 121–135 (127.3 ⫾ 4.5), F (n ⫽ 14) 119–141 (128.5 ⫾ 6.5); tail, M 118–147 (130.4 ⫾ 7.9) (the largest, with wing 135, tail 147, and wing 135, tail 144, were resident in May and August), F 104–140 (120.6 ⫾ 9.0), bill, M 16.8–20.2 (18.3 ⫾ 1.2), F 17.4–19.7 (18.2 ⫾ 0.6), U (n ⫽ 7) 16.0–19.4 (17.6 ⫾ 1.4) (AMNH, BMNH, FMNH, MCZ, RMNH,YPM); S. l. brachyurus, “Malacca,” peninsular Thailand and Malay Peninsula, April-September, wing, M (n ⫽ 16) 96–127 (121.6 ⫾ 3.7), F (n ⫽ 8) 113– 126 (120.9 ⫾ 4.4); tail, M 98–120 (113.8 ⫾ 5.2), F 104–125 (16.2 ⫾ 7.2) (ANSP, USNM, ZRC); Borneo, April-September, wing, M (n ⫽ 14) 117–128 (120.6 ⫾ 3.4), F (n ⫽ 11) 112–132 (119.9 ⫾ 6.3); tail, M 106–128 (113.8 ⫾ 7.3), F 108–129 (116.4 ⫾ 7.5) (AMNH, MCZ, MNHN, MZB, RMNH, ROM,WFVZ, ZMB, ZRC); Sumatra; April–September, wing, M (n ⫽ 16) 116–129 (122.1 ⫾ 4.1); F (n ⫽ 18) 117–130 (122.2 ⫾ 8.7); tail, M 112–136 (111.5 ⫾ 29.8), F 98–130 (115.8 ⫾ 9.2) (ANSP, MZB, RMNH, USNM, ZMA, ZRC); Palawan and Balabac: Wing, M (n ⫽ 20) 117– 127 (122.6 ⫾ 3.2), F (n ⫽ 28) 92–128 (123.3 ⫾ 2.4); tail, M 96–127 (111.2 ⫾ 10.6), F 97–127 (110.0 ⫾ 9.3) (AMNH, ANSP, BMNH, BPBM, CM, DMNH, MNHN, RMNH, ROM, USNM, ZRC). Weight, Borneo, M (n ⫽ 6) 26–31 (29.7), F (laying?) (n ⫽ 3) 32–36 (34) (MZB). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⱖ 1 0 ⬎ 2 ⬎ 1; in other birds, . . . 2 ⬎ 1 ⬎ 10.

Field characters Overall length 24 cm. Black bird with white pants. Slender black cuckoo with tail broad at tip and tail slightly forked to square, the under tail coverts with whitish bars. Differs from northern migrants that winter in its range, in its square tail (forked tail in Fork-tailed Drongo-cuckoo S. dicruroides). Differs from drongos in its white leggings, white bars on under tail coverts and inconspicuous white bars on the tail, thin decurved bill, black iris (red in drongos), and a white bar on underside of the wing in flight.

Voice Song, loud clear mellow whistles rising evenly up the scale,“one, two, three, four, five, six” (“as if some one were practicing a musical scale in the wilds of the jungle”, Legge 1880). In contrast, in song of Fork-tailed Drongo-cuckoos in the Himalayas the second note is lower or at the same pitch as the first note, and the sound is more piping.A song with the Square-tailed Drongo-cuckoo arpeggio “one, two, three . . .” pattern at Kaziranga,Assam (NSA 15022, Figure 6.3 g), indicates that this species occurs in mainland southern Asia north to the Brahmaputra River. Songs are consistent across its range in southeast Asia and the Malay Archipelago (Scharringa 1999, Ball 2002, Supari 2003, Figure 6.3). The cuckoo also has a long “crescendo” call, a series of burry or buzzy rising notes ending with descending notes of the same harsh tonal quality, given in excitement as do other drongocuckoos.

Range and status In India in Assam (Kaziranga NP and Nameri NP where identified by song, and birds with unforked tails and broad rectrices occur in the breeding season in the Garo Hills, Lushai Hills and Khasi Hills). In southeast Asia they occur in Burma (Arakan, Pegu, Shan States), Thailand, the Malay Peninsula, Sri Lanka, Sumatra, Java, Bali, Bangka, Borneo, Palawan, Balabac and Çalauit.Ali (1953) noted them in “practically all India, patchily; also Assam” with fork-tailed birds in the Himalayas, then later (Ali 1969) considered all drongo-cuckoos in India to be the forktailed birds. Resident in most of the range, seasonal in Kaziranga, Assam (Barua and Sharma 1999).

462 Square-tailed Drongo-cuckoo Surniculus lugubris

? ?

saplings in burned clearings. They feed in the canopy, and they sally into swarms of flying termites and hawk insects like a drongo (Wells 1999, Strange 2001).The flight is smooth and direct, not dipping and acrobatic like that of a drongo.

Food Mainly insects, caterpillars (both hairy and hairless: Pieridae, Lasiocampidae, Limadodidae), beetles, other soft insects, spiders; fruit, mainly figs (Legge 1880,Thompson 1966, Smythies 1981, Sody 1989).

Displays and breeding behavior Traced by song these cuckoos occur in India (Assam: Kaziranga NP, NSA 15022),Thailand (west, Huai Kha Khaeng; peninsular,Thaleban NP, both P. Round recordings; Khao Phraa Bang Kran, NSA 41320); Sri Lanka (Deepal Warakagoda recordings); Peninsular Malaysia (Ampang, NSA 31616, 34921; Pasoh Forest Reserve, NSA 07588, 34964; Taman Negara NP, NSA 06259, 06260, 07588; Gombak, NSA 32616), Sumatra (Way Kambas Reserve, Lampung, NSA 06806), Western Kalimantan (Cabang Panti, Matan, J. M. Lammertink recordings), Sabah (Danum Valley, NSA 36035) and Palawan (Mt St Paul’s NP, NSA 17837).

Habitat and general habits Mainly hill forest, also on the plains.They occur in forests and scrub, in foliage canopy, in orchards of cashew, jackfruit and coconut and in pepper vine plantations in southern India, interior of dry forests in Sri Lanka, lowland semi-evergreen and evergreen forest and in tree plantations, pines in the Malay Peninsula, and in recently logged primary forest in Sabah (Deignan 1945,Wells 1999, Strange 2000, WFVZ). In Assam they occur in semi-evergreen forest and in swamp forest and riparian woodland in lowlands along the Brahmaputra River and in nearby hills (Barua and Sharma 1999, RBP, FMNH, UMMZ, NSA).The altitudinal range in Malay Peninsula is from the lowlands to 1200 m (Wells 1999), in Sri Lanka from lowlands to 1300 m (Harrison 1999), in Java and Borneo to 1200 m (Hoogerwerf 1948, Smythies 1957, Sheldon et al. 2001). Arboreal, they perch on charred stumps and

Males are territorial, they call loudly back and forth from exposed tree perches (Wells 1999, RBP).

Breeding In Kerala they breed from January to March (this species?) and in Sri Lanka from December to May, based on song, enlarged gonads and breeding of hosts (Ali and Ripley 1969, Kotagama and Fernando 1994, BMNH). In the Malay Peninsula they sing from January to August and fledglings appear from May to July and an egg was found in July (Wells 1999); in Java eggs are found from January to May and in July, September, November and December and the cuckoos sing nearly all year (Hoogerwerf 1949), while in Borneo they sing from December through March and females had oviduct eggs in April and May (Smythies 1957, Sheldon et al. 2001, WFVZ). Resident birds in the Malay Peninsula are in wing molt from July to September after their breeding season (Wells 1999). Brood-parasitic.The hosts are mainly babblers. In India the cuckoos use small babblers, Nepal Fulvetta Alcippe nipalensis and Dark-fronted Babbler Rhopocichla atriceps, and forktails Enicurus, bulbuls Pycnonotus aurigaster, shrikes Lanius species and Striated Warbler Megalurus palustris (Baker 1906, 1934, Ali and Ripley 1969, Becking 1981). In Sri Lanka they use Darkfronted Babbler Rhopocichla atriceps (Becking 1981; BMNH 1924.9.5.2), in the Malay Peninsula they parasitize small tit-babblers: Striped Tit-babbler Macronous gularis (these mob the cuckoo, Duckworth 1997), Sooty-capped Babbler Malacopteron affine, Chestnut-winged Babbler Stachyris erythroptera and Horsfield’s Babbler Malacocincla (Trichastoma) sepiarium

Moluccan Drongo-cuckoo Surniculus musschenbroeki 463 (the first three babblers with cuckoo fledglings fed by pairs, the last with an egg perhaps of this cuckoo, Wells 1999); in Java the hosts are mainly Malacocincla sepiarium, also Grey-cheeked Tit-babbler Macronous flavicollis and Chestnut-chested Babbler Stachyris melanothorax (Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967, Becking 1981); and in Borneo Chestnut-winged Babbler S. erythroptera (Cranbrook and Wells 1981).There are no recent observations of drongos as hosts, although Makatsch (1955) listed Black Drongo Dicrurus macrocercus. Some early records involved young cuckoos in the care of drongos. Jerdon (1862) saw a Black Drongo in Burma chase a drongo-cuckoo (identified by white in the plumage), suggesting the drongo was a host and the cuckoo was

near its nest;T. R. Bell (in Baker 1906) saw a young drongo-cuckoo fed by Black Drongo, and Davison saw a young drongo-cuckoo with a pair of Crowbilled Drongos D. annectans, and the cuckoo was fed by a drongo (Hume and Davison 1878). In addition, Lewis (1898) saw drongos tormenting a cuckoo and pecking it to death, not a case of successful parental care. The egg in Borneo is white with fine purple splotches (Smythies 1957), in Java pale salmon-pink with brownish, purplish or gray markings. Eggs in the Malay Peninsula in the nest of Horsfield’s Babbler M. sepiarium, 22.5 ⫻ 17 mm (Wells 1999); eggs in Java 20 ⫻ 15.5 mm (Hoogerwerf 1949), an egg in oviduct in Sabah 19x14 (WFVZ 37595, Sheldon et al. 2001). The incubation and nestling periods are unknown.

Moluccan Drongo-cuckoo Surniculus musschenbroeki A. B. Meyer, 1878 Surniculus musschenbroeki A. B. Meyer, 1878, Rowley’s Ornithological Miscellany, 3, p. 164. (Bacan, Moluccas).

SOURCES: AMNH, ANSP, BMNH, MZB, RMNH, SMF, USNM, ZMB, ZMS.

Description

Wing, M (n ⫽ 7) 125–136 (131.1 ⫾ 4.2), F (n ⫽ 5) 122–135 (128.6); tail, M 121–139 (130.5 ⫾ 7.8), F 116–137 (121.5), bill, M (n ⫽ 2) 18.7–19.2 (19.0), F 18.3–21.4 (20.2 ⫾ 1.1); tarsus, M (n ⫽ 2) 15.5–16.4 (15.8), F 15.4–17.5 (16.3 ⫾ 0.8) (AMNH, BMNH, MZB, RMNH, SMF, USNM, ZMB). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⱖ 10.

ADULT: Sexes alike, head, upper back and breast velvety purplish black, nape with small concealed white patch, wing black with purplish, blue or green gloss, tail black with purplish, blue or green gloss, tip black, tail feathers T3, T4 broad near the tip and often curled upwards and outwards, the tail square to slightly forked, the outer tail feathers T5 shorter and with narrow oblique white bars and white spots near the shaft at the base of the feather; underparts black, leggings white (long feathers on leg, outer vane of feather white, inner vane black or white), under tail coverts black (not barred), under wing coverts black, white bar across wing base inner webs of P1–7 and S1–4 seen from below, P8 unmarked black, P9 with white spot near the base of inner web; skin around eye dark bluish gray, iris brown, bill black, feet dark gray. JUVENILE: Plumage dull black with white spots on the crown, wing coverts, throat, breast and under tail coverts (USNM 571628, SMF 28001). NESTLING: Undescribed.

Measurements and weights

Systematic comments These birds are distinct in song from other drongo cuckoos. The adult plumage is velvety, not glossy; the tail is longer and more recurved than in Philippine Drongo-cuckoos S. velutinus. The juvenile plumage is dull black with white spots, as in the two Asian drongo-cuckoos. As recognized by Salvadori (1881), these birds are a distinct species.

Field characters Overall length 23 cm. Black bird with white pants. A black arboreal cuckoo with contrasting head and underparts dull velvet and upperparts glossy, tail square, white leggings and black under tail coverts. The velvety body plumage lacks the metallic green

464 Genus Hierococcyx color of Fork-tailed Drongo-cuckoo S. lugubris and the wing and tail are not as glossy as those of Philippine Drongo-cuckoo S. velutinus. In flight, white bar on underside of wing. Looks like a drongo, but the bars in tail and the fine bill distinguish it as a cuckoo.

Voice Song a series of eight or so whistles, the series of successive notes rising only slightly in pitch, each note falling in pitch. On Sulawesi, the song is “ki ki ki ki ki ki ki ki” with 8–12 notes nearly equal in pitch in a series.The song lasts about 2 sec with a pause of 2 + sec before the next song (Heinrich in Stresemann 1940, Watling 1983, Holmes and Phillips 1996, Coates and Bishop 1997). On Halmahera “the call had a slightly different quality to Malaysian birds [⫽ Square-tailed Drongo-cuckoo S. lugubris], with a series of 6–8 notes, and occasionally as many as 12” (Lambert and Yong 1989) (Figure 6.3).

Range and status Sulawesi, Buton and the northern Moluccas (Bacan, Halmahera). In Sulawesi, it is uncommon or rare (Stresemann 1940, van Bemmel and Voous 1951, White and Bruce 1986, Coates and Bishop 1997). In the Moluccas, musschenbroeki is known from Bacan (⫽ Batjan, Batcian) by three specimens: (1) Meyer’s holotype (formerly in Dresden Museum, lost in the war, S. Eck, in litt.), (2) SMF 27996, described by von Berlepsch (1901), and (3) Braunschweig Staatliches Naturhistorische Museum, taken by Platen (Nehrkorn 1894); for Halmahera it is known from specimens in SMF and USNM. Salvadori (1881) questioned the Bacan record, but Kükenthal’s specimen has a date on the label (SMF 27996, 1 June 1894).

Habitat and general habits Primary forest; in Sulawesi at 300 m and 500 m to 1200 m, on Halmahera at 250 m, on Bacan 100– 200 m, and on Obi at 360 m (Stresemann 1940, Ripley and Rabor 1958, Lambert and Yong 1989, Lambert 1994, Coates and Bishop 1997). One report says they are nocturnal in behavior (Riley 1924), perhaps calling at night.

Breeding Song was noted in northern Sulawesi in March, April and September, and a pair displayed in October (Rozendaal and Dekker 1989, Wardill et al. 1998). Presumably brood-parasitic, the hosts are unknown. Heinrich (in Stresemann 1940) saw two Hair-crested Drongos Dicrurus hottentottus call and chase a calling drongo-cuckoo, but drongos are aggressive and chase many kinds of birds, and a chase may not indicate a host-parasite interaction. Eggs and other details of breeding are unknown.

Genus Hierococcyx S. Müller, 1845 Hierococcyx S. Müller, 1845, Verhandlingen over de Natuurlijke Geschiedenis der Nederlandsche overzeesche zebittingen, Land-en Volkenkunde, pt. 8, p. 233. Type, by original designation, Cuculus fugax Horsfield. Large to medium-sized cuckoos with broad to pointed wings and a broadly barred tail. Cabanis and Heine (1863) substituted the name Hiracococcyx, for etymological

reasons, suggesting that Müller as a parsimonious Hollander avoided a six-syllable word even though the longer word was the correct application of “hawk-cuckoo” (the word denotes “hawk-cuckoo”, not “holy cuckoo”). Nevertheless, etymological correctness is not an admissible argument for a change in names, according to nomenclatorial codes (e.g.

Mustached Hawk-cuckoo Hierococcyx vagans 465 ICZN 1999). Peters (1940) disregarded the genus Hierococcyx because cuckoos with these characters were not exclusive from other cuckoos that were not considered members of the same set, and the wing shape varied within the set. The secondaries are nearly as long as the primaries, nearly 2/3 the length

of the primaries in the folded wing; in contrast to Cuculus where the primaries are longer.The genetic results show that Hierococcyx cuckoos in the sense of H. fugax and allied species, H. sparverioides and allied species and H. vagans together form a monophyletic set. Eight species.

Mustached Hawk-cuckoo Hierococcyx vagans (S. Müller, 1845) Cuculus vagans S. Müller, 1845, Verhandelingen over de Natuurlijke Geschiedenis der Nederlandsche overzeesche bezittingen, Land-en Volkenkunde, 8, p. 233. ( Java) Monotypic. Other scientific names: Cuculus nanus Hume, 1877.

Description ADULT: Sexes alike, crown dark gray to black, back dark brown, wing dark brown barred lighter brown, tail with broad bars of light gray and black with tip edged white, face with broad black line extending down and back from in front of the eye, another black line extends down and back from behind the eye, cheek and ear patch white, collar brown; underparts white with fine and distinct dark streaks, under tail coverts white, unbarred, bend of wing white, under wing coverts unmarked whitish, primaries from below mainly whitish with gray bars; eye-ring yellow, iris dark brown to gray, bill black, base greenish, below yellow-green, drooping at tip, feet yellow. SUBADULT: Resembles adult but has obscure gray and rufous bars on the dark brown upperparts, gray edges of feathers on rump and upper tail coverts and pale, rufous-tipped flight feathers, perhaps retained from juvenile plumage (ZRC 3.7933). JUVENILE: Crown blackish, back blackish with obscure rufous bars, wing blackish with distinct pale rufous bars and tips of flight feathers, tail gray barred black as in adult, with white tips, face blackish with indistinct broad black line in front of eye continuous with blackish throat, the throat and chin blackish, breast with thick black streaks under tail coverts unmarked white, iris dark brown (ZRC 3.7931). NESTLING: Undescribed.

SOURCES: AMNH, BMNH, MCZ, MZB, RMNH, SMTD, USNM,YPM, ZMB, ZRC.

Measurements and weights Wing, M (n ⫽ 13) 135–156 (145.0 ⫾ 5.2), F (n ⫽ 10) 136–150 (143.1 ⫾ 4.9); tail, M 120–144 (131.4 ⫾ 7.5), F 113–145 (125.2 ⫾ 10.5); bill, M 19.3–21.7 (19.8 ⫾ 0.8), F 19.6–22.6 (21.5 ⫾ 1.0); tarsus, M 17.2–23.1 (19.2 ⫾ 0.8), F 17.2–20.2 (18.6 ⫾ 1.2) (AMNH, BMNH, MCZ, MZB, USNM,YPM, ZRC). Weight, M (n ⫽ 1) 54.5–63.4 (58.2) (repeat weights of a ringed bird, Wells 1999), F (n ⫽ 1) 62 (BMNH), U (n ⫽ 1) 53 (Wilkinson et al. 1991). Wing formula, P7 ⫽ 6 ⬎ 8 ⬎ 5 ⬎ 4 ⬎ 9 ⬎ 3 ⬎ 2 ⬎ 1 ⬎ 10.

Field characters Overall length 26 cm. Small cuckoo, streaked below, with a distinct face pattern like a small Peregrine Falco peregrinus. In flight, the inner webs of inner flight feathers and coverts are whitish and form a pale patch when the wing is seen from below.

Voice A two-syllable phrase “chu-chu”, repeated every 2 sec, the two notes descending from 1.8 to 1.4 kHz, and the phrase lasting 0.4 sec. Other call is a mellow whistle “peu peu” the notes first given singly, then in hurried couplets ascending on scale in a crescendo, then a sudden stop,“hee hee hee, . . . hi-hi hi-hi hi-hi hi-hi” with short notes rising 1.3 to 1.5 kHz, given at rate of 8 notes in 2 sec in a trill lasting 6–8 sec (Smythies 1981, Thomas and Thomas 1994, Scharringa 1999, Wells 1999, Robson 2000a, NSA 15537).

466 Dark Hawk-cuckoo Hierococcyx bocki

Range and status Southern Burma (Tenasserim, Mt Nwalabo, Mergui Archipelago), southern Thailand (Chon Buri in southeast, peninsula from Chumphon to Phuket) and southern Laos (Xe Piane) south through Malay Peninsula to Sumatra, Java and Borneo, including Sarawak (Gunung Mulu NP), Sabah and Kalimantan. Resident. In the Malay Peninsula, Gibson-Hill (1949b) noted the species in September and October and thought it was a migrant; Wells (1999) found birds in song, on territories and breeding, and one was caught three times over 13 months, indicating a resident status. Uncommon in southeast Asia, rare in Sumatra and Borneo, no recent records in Java (Kuroda 1936, Smythies 1986, 1999, Deignan 1963, Cranbrook and Wells 1981, Smythies 1986, Round 1988, van Marle and Voous 1988, Wilkinson et al. 1991, MacKinnon and Phillipps 1993, Duckworth 1996, Thewlis et al. 1996, Holmes 1997, Round and Treesucon 1997, Robson 2000a, Sheldon et al. 2001). Conservation status, near-threatened (BirdLife International 2001).

Habitat and general habits Forest edge, primary and secondary forest, riverine and alluvial forest, bamboo forest. They live in the understory and lower canopy of dry lowland forest. In Laos, seen in semi-evergreen forest dominated by bamboo. They occur in the lowlands and on lower hills to 900 m in the Malay Peninsula, from sea level to 550 m in Sabah, and at 130–150 m in central Kalimantan. Males are territorial (Wilkinson et al. 1991,Wells 1999, Sheldon et al. 2001).

Food Presumably large insects.

Breeding In Malay Peninsula the birds peak in song in February and young appear in July and August (Medway and Wells 1976, Cranbrook and Wells 1981, Wells 1999); in Aceh, northern Sumatra, birds sing in August (van Marle and Voous 1988); in Kalimantan the birds sing in March (Holmes 1997). Broodparasitic. In peninsular Thailand at Krabi, a fledgling with mustached face and a black and white tipped tail was attended and fed by Abbott’s Babbler Malacocincla (Trichastoma) abbotti (P. Round, in Wells 1999). In Sabah, Borneo, a nestling perhaps of this cuckoo species was fed by Rufous-winged Philentoma Philentoma pyrhopterum (Sheldon et al. 2001). The eggs and other details of breeding are unknown.

Dark Hawk-cuckoo Hierococcyx bocki Wardlaw Ramsay, 1886 Hierococcyx bocki Wardlaw Ramsay, 1886, Ibis, 1886, 157. (mountains of W. Sumatra) Monotypic.

Description ADULT: Sexes alike, crown, face and neck dark gray (no white in front of the eye), back and rump slate gray, wing unmarked brown, inner vane of flight

feathers with incomplete white bars, tail long, broadly barred gray and narrowly barred black edged with orange-buff on distal side of each black bar, the terminal black bar broad and tail tipped with whitish; face uniformly dark (not pale in front of eye); underparts with chin and throat gray, breast dark rufous, belly and flanks white with black bars, under tail coverts unbarred white, bend of wing

Dark Hawk-cuckoo Hierococcyx bocki 467 white, under wing coverts white barred dark gray and the marginal coverts rufous; eye-ring yellow, iris orange to buff, bill blackish above and green below, feet yellow. SUBADULT: Crown uniform chocolate brown, back slate with indistinct brown bars, wings brown with rufous notches on the outer vane, throat white, breast white broadly barred and splotched dark brown, belly and under tail coverts white (MCZ 197760).

and in wintering mainland hawk-cuckoos H. fugax, H. nisicolor and H. hyperythrus). In wintering area differs from Large Hawk-cuckoo H. sparverioides by darker plumage, dark face, small size, and song (H. bocki sings in season of overlap; H. sparverioides is silent and when it does sing the song differs); subadult is barred not streaked below and lacks rufous on breast. Juvenile differs from other hawk-cuckoos in its range in the uniformly brown upperparts and the brown chin and breast.

Voice JUVENILE: Crown unmarked slaty brown with a few white feathers on nape, the back and rump dark brownish black, tail broadly banded as in adult, underparts with chin unmarked slaty brown, breast slaty brown with white base of feathers, lower breast, belly and under tail coverts white with dark bars, wing dark brown with indistinct rufous notches on outer edge of primaries and buff tips, the secondaries uniform gray brown, bend of wing white (AMNH 625573, MZB 17963).

Malay and Borneo birds interpreted as “pee-ha, peeha” by Wells (1999), “pit-piwit” by T. Harrison (AMNH 648292),“pi-phu” or “CHUPCHUPchee” (Robson 2000a), not “brain fé-ver” as in the Large Hawk-cuckoo H. sparverioides (van Balen 1997, Davison 1997, Robson 2000a). The song is a twonote whistle; the first note rises then falls from 1.7 to 2 to 1.7 kHz, and the second note rises then falls gradually with a catch near the end, beginning at 1.7 and rising to 1.9 or 2.0 then down to 1.7 kHz, without a burry tremolo (Figure 6.2).

NESTLING: Undescribed.

Range and status SOURCES: AMNH, BMNH, MCZ, MZB, RMNH, SMTD, ZMA, ZMB, ZRC.

Borneo and Sumatra, wing, M (n ⫽ 16) 184–196 (186.3 ⫾ 5.9), F (n ⫽ 7) 180–200 (189.7 ⫾ 5.1); tail, M 132–175 (151.0 ⫾ 13.2), F 136–175 (155.0 ⫾ 3.2); bill, M 18.4–21.6 (21.1 ⫾ 11.2), F 19.5–23.2 (21.1); tarsus, M 18.6–23.2 (20.3 ⫾ 2.0), F 19.2– 21.4 (19.9) (AMNH, BMNH, MCZ, MZB, RMNH, SMTD, ZMA, ZMB, ZRC). Weight, U (n ⫽ 2) 130–131 (130.5) (Wells 1999). Wing formula, P8 ⬎ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⫽ 10.

Malay Peninsula, Sumatra and Borneo. Resident, mainly in the hills and mountains. In Malay Peninsula, they are common but local residents. In Sumatra, resident Dark Hawk-cuckoos are in submontane and lower montane forests at Utara (Bandarbaru, Berastagi, Mt Sibayak, Batak Highlands), Barat (Mt Sago, Padang Highlands) and Bengkulu (Redjang) (van Marle and Voous 1988). In Borneo, residents are locally common in mountains above 1000 m with most observations on Kelabit Plateau (Tamabo Range) in Sarawak and on Mt Kinabalu and Mt Trus Madi in Sabah (Smythies 1981, van Balen 1997, Davison 1997, Sheldon et al. 2001).

Field characters

Habitat and general habits

Overall length 30–32 cm. Cuckoo with an Accipiter hawk-like appearance, nearly black above, chin and throat ashy gray, breast rufous, lower breast and belly barred black and white, wings short and tail long and banded, the black bands as broad as the gray bands (black bands narrower in other resident

Wooded areas, in open and closed forest. In the Malay Peninsula, they occur between 1100– 1800 m, in forest edge, tall secondary forest and open woodland (Wells 1999). In Sumatra they occur in pine plantations at 900 to 1600 m (van Marle and Voous 1988). They hawk flying

Measurements and weights

468 Large Hawk-cuckoo Hierococcyx sparverioides Kooiman 1951,Wells 1999). Arboreal and solitary, their behavior has not been observed in detail. Males hold territories and countersing with a neighbor (Wells 1999).

Food Insects, mainly caterpillars, and crickets, grasshoppers, beetles, bugs, roaches, ants and spiders, bird eggs, and berries (Smythies 1981, Sody 1989).

Breeding termites from tops of trees, and they fly from a perch to the ground for food ( Junge and

They sing September to late May in the Malay Peninsula (Wells 1999). Presumably broodparasitic. The eggs and other breeding details are unknown.

Large Hawk-cuckoo Hierococcyx sparverioides (Vigors, 1832) Cuculus sparverioïdes Vigors, 1832, Proceedings of the Committee of Science and Correspondence of the Zoological Society of London 1, 173. (Himalayas) Monotypic.

unbarred new feathers and old barred flight feathers with rufous bars on outer vane and the inner part of inner vane, throat and breast rufous with blackish streaks.

Description

JUVENILE: Forehead and crown blackish with a few white feathers, nape dark brown, feathers rufous at base, back and rump dark brown with light rufous bars, broad dark bands in the tail, below the chin black, underparts white with bold black streaks on breast, splotched on belly and barred on flanks, under tail coverts white with incomplete black bars, wing dark brown with light rufous bars and notches on outer edge of primaries, bend of wing white; bill black.

ADULT: Sexes alike, crown, face and neck dark gray, back and rump brown, wing unmarked brown, inner vane of flight feathers with incomplete white bars, tail long, broadly barred gray and narrowly barred black edged with orange-buff on distal side of each black bar, the terminal black bar broad and tail tipped with whitish; face paler in front of eye; underparts with chin blackish, throat white obscured with gray or rufous or streaked blackish, breast rufous with brown streaks, belly and flanks white with black bars, under tail coverts unbarred white, bend of wing white, under wing coverts white barred dark brown; eye-ring yellow, iris orange to buff, bill blackish above and green below, feet yellow. SUBADULT: Crown blackish brown (not gray as in adult), back and rump brown indistinctly barred rufous, wing barred rufous, underparts less rufous than in adult plumage. Some birds retain part of the juvenile plumage in first adult year, wing with both

NESTLING: Undescribed. SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MVZ, MZB, RMNH, ROM, UMMZ, USNM, ZMA, ZRC.

Measurements and weights India: Wing, M (n ⫽ 11) 227–246 (2 3 6 ⫾ 5.50), F (n ⫽ 9) 210–243 (226.7 ⫾ 11.1); tail, M 210–243 (223.4 ⫾ 11.3), F 195–228 (2 1 3 ⫾ 10.6); bill, M 24.0–26.7 (25.1 ⫾ 0.9), F 23.3–26.7 (24.3 ⫾ 1.1);

Large Hawk-cuckoo Hierococcyx sparverioides 469 tarsus, M 24.0–29.4 (26.4 ⫾ 1.5), F 22.1–24.8 (23.4 ⫾ 0.9) (UMMZ). Weight, M (n ⫽ 3) 140–160 (150.7) (AMNH, UMMZ). Wing formula, P8 ⬎ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1 ⫽ 10.

Field characters Overall length 38–40 cm. Large cuckoo with an Accipiter hawk-like appearance, gray head contrasts with brown back, wings short and tail long and banded, the black bands as broad as the gray bands (black bands narrower in other mainland hawkcuckoos H. fugax, H. nisicolor and H. hyperythrus, broad in Sulawesi Cuckoo Cuculus crassirostris with black and rufous bands); differs from Common Hawk-cuckoo H. varius by the blackish chin and in subadult plumage the underparts more boldly streaked and barred; in winter from Dark Hawkcuckoo H. bocki by paler plumage, light face and large size. Subadult has less brown on breast than subadult H. bocki. Juvenile differs from other hawkcuckoos in its range by the upperparts being brown barred rufous, underparts with bold black streaks on throat and breast and bold bars on belly and flanks (in C. bocki, crown and back unmarked brown, underparts without rufous, chin unmarked brown, breast brown with white base of feathers, rest of underparts white without dark bars). Juvenile differs from other hawk-cuckoos in its range by the bold black streaks on throat and breast and bold bars on belly and flanks.

Voice Shrill, screaming two-note whistle,“pi-peeah”, the first note short, beginning above 2.2 kHz and dropping in pitch, and the second note long sliding from low to high (1.8 to 2.4 kHz) then dropping rapidly to end in a gurgle, vibrato or tremolo at 2.1 kHz.When the bird is excited the phrases are repeated in runs of 5 or 6, increasing in loudness and pitch (the first note begins at 3 kHz, the second at 2 kHz rising to 2.8 kHz, and so on for later phrases) (Ali 1946, 1962, 1996, White 1984, Martens and Eck 1995, Scharringa 1999, P.Alström tapes, P. Holt tapes; Figure 6.2). The call has also

been interpreted as “brain fé-ver” (Smythies 1940, 1960) leading to misidentification with Common Hawk-cuckoo H. varius, the other “Brain-fever Bird”. Song is nearly identical from male to male, a loud and far-carrying sound in the oak zone of the Himalayas. Another call, a crescendo series of harsh notes rising then falling in pitch, given when two males sing close together and when a male responds to playback of his own song (Arunachal Pradesh: RBP). Female call, a series of wavering whistles “droo droo droo” (Martens and Eck 1995). Birds call in the mornings, at sunset and into the night.

Range and status Mainland Asia in the Himalayas to hills of southeast Asia. They occur in the hills in northern Pakistan (Murree, Dunga Gali, Ghora Gali) where rare and not seen in recent years (Roberts 1991), in India from Himachal Pradesh east to Arunachal Pradesh, Naga Hills, Manipur and Lushai Hills, and Nepal and Bhutan (Rand and Fleming 1957, Grimmett et al. 1999, UMMZ), China from southern Sichuan and Yunnan to the lower Yangtse valley and Beijing, Guangsi and Hainan (Guangdon et al. 1983, Cheng 1991, Vuilleumier 1993) and southeast Asia in Burma (Chin Hills, Shan States, northern Burma), Thailand and Indochina (Smythies 1986, Lekagul and Round 1991, Robson et al. 1998, MacKinnon and Phillipps 2000, King et al. 2001). In winter they occur in the north along the base of the eastern Himalayas and in southern India in the Eastern and Western Ghats,

470 Large Hawk-cuckoo Hierococcyx sparverioides Bangladesh (Thompson et al. 1993), Nicobar Is (Sivakumar and Sankaran 2002), Malay Peninsula, Sumatra, Java, Bali, Borneo, Sulawesi and the Philippines. Migratory in the north of its range and resident in the south. There are a few winter records in Yunnan and the Himalayas, most birds move south to winter; they are resident in Assam (UMMZ) and sing in November in northernmost Burma (King et al. 2001). In the Malay Peninsula migrants occur from late October to late March. They move at night, with migrants or winter visitors appearing in mist nets at a mountain pass (Fraser’s Hill), and one was attracted to lights in November; cuckoos appear in passage in midNovember at Singapore (Medway and Wells 1976, Wells 1999). In the Greater Sunda Islands they occur in winter. In Sumatra, birds occur in Utara and Lampung (van Marle and Voous 1988). In western Java, birds are uncommon from November to March (Kuroda 1936, RMNH). In Borneo, wintering migrants are occasional in the north (Labuan, Kuching) (Smythies 1981, van Balen 1997, Davison 1997). In northern Sulawesi they are rare winter visitors (White and Bruce 1986). In the Philippines they appear from late September to May and are seen mainly in the north (Busanga, Luzon, Negros, Palawan, Mindoro, Leyte and Polillo) (Kennedy et al. 2000). Breeding density, in northern India as many as three birds per square mile (c. 1 bird per km2) (Fleming 1967).

Habitat and general habits Wooded areas, in open and closed forest. Occur in oaks and firs in cloud forest in Himalayas (Fleming 1967, Inskipp and Inskipp 1985, Martens and Eck 1995); breed from 900 to 3000 m in the Indian subcontinent (Fleming 1967, Grimmett et al. 1999), and seen as high as 2135 m in southeast Asia (Robson et al. 1998) not known to occur together with Common Hawk-cuckoo H. varius, which live in the lowland plains. In Thailand in deciduous and evergreen woodland from foothills to 2500 m (Lekagul and Round 1991). In Malay Peninsula passage and wintering H. sparverioides are in forest edge, tall secondary forest, open woodland, gardens, thickets and mangroves, mainly in low country (Wells 1999). In the Philippines wintering birds occur in

trees in open grassland (Kennedy et al. 2000). In Borneo wintering birds are in the lowlands, while resident Dark Hawk-cuckoo H. bocki are in the mountains above 800 to 1000 m (Smythies 1981, Sheldon et al. 2001). Arboreal, solitary, they remain in the tree canopy. In flight they resemble a small hawk Accipiter, passing low with a few fast wingbeats then a glide, and rising at the end of a glide to perch in a tree. Large Hawk-cuckoo Hierococcyx sparverioides as its descriptive specific name indicates looks much like Eurasian Sparrowhawk Accipiter nisus, in wing and tail shape and broad body as well as in plumage pattern and color. A Large Hawkcuckoo we observed at Chandra, Arunachal Pradesh, India, sang, and when song was played back the cuckoo approached at once, glided directly toward the speaker, and in flight it looked very much like a sparrowhawk. On repeated approaches the bird flapped and glided, like a sparrowhawk. We photographed the hawkcuckoo as it repeated its performance; the Figure opposite shows the similarity of the two profiles in flight, the hawk-cuckoo traced from our photographs and the sparrowhawk from photographs in Porter et al. (1976). The bird appeared hawklike not only to us: small tits and warblers near us dove into the bushes and gave high-pitched “see” notes, as they do to a hawk. Large Hawk-cuckoos and Eurasian Sparrowhawks occur together in this region. It is unknown whether their plumage and behavior aids the hawk-cuckoos in parasitism when they lay an egg in the nest of their laughingthrush hosts.

Food Insects, mainly caterpillars and crickets, grasshoppers and beetles, bugs, roaches, ants and spiders, bird eggs, and berries (Smythies 1981, Sody 1989).

Breeding In hills of western India they first sing in March and April, then at night in June and July, the main laying season ( Jerdon 1862, Hume 1873). In Nepal they sing from March to July or August, in Sikkim a female had an oviduct egg in June (Baker 1906, Ali 1962), in Assam they sing and breed from April

Common Hawk-cuckoo Hierococcyx varius 471 Cuckoo

Hawk

Flight patterns of Large Hawk-cuckoo Hierococcyx sparverioides and Eurasian Sparrowhawk Accipiter nisus (hawk from Porter et al. 1976).

to June (Ali and Ripley 1969), in Burma they call and lay from February to June (Mackenzie 1918, Smythies 1940); in Pakistan there are no breeding records (Roberts 1991).

Brood-parasitic. Hosts, in India they parasitize Streaked Laughingthrush Garrulax lineatus, Chestnut-crowned Laughingthrush G. erythrocephalus, Himalayan Whistling-thrush Myiophonus caeruleus, Nepal Shortwing Brachypteryx leucophrys, Hoary Barwing Actinodura nipalensis, Large Spiderhunter Arachnothera magna and other songbirds (Hume 1873, Baker 1906, 1934, Becking 1981); in Burma they use Lesser Necklaced Laughingthrush G. monileger (Mackenzie 1918). Eggs are of two morphs, (1) light brownish olive, seldom with darker specks, and (2) blue, unspotted (Baker 1906, 1907, Osmaston 1912, 1916); 27 ⫻ 19 mm (Baker). Becking (1981) agreed with Baker’s identifications and host records of brown eggs in nests of spiderhunters and shortwings, but doubted the identity of blue eggs in nests of babblers and thrushes and argued these were eggs of Common Cuckoo Cuculus canorus. Becking reported the ultrastructure of these eggshells to be like that of C. canorus, but he did not report the ultrastructure of eggs of H. sparverioides or Common Hawk-cuckoo H. varius and these may be similar to C. canorus. He stated that the blue eggs were the same size as the eggs of C. canorus, but Osmaston’s (1912, 1916) measurements of four blue eggs were 27.9–30.0 ⫻ 20.6–21.3 mm, all larger than the eggs of C. canorus (23 ⫻ 17 mm). Eggs are the same color and length and less rounded than the blue eggs of Chestnut-winged Cuckoo Clamator coromandus.The incubation and nestling periods are unknown.

Common Hawk-cuckoo Hierococcyx varius (Vahl, 1797) Cuculus varius Vahl, 1797, Skrivter af Naturhistorie-Selskabet, Kjøbenhavn , 4, Heft 1, p. 60.Tranquebar [India] Polytypic.Two subspecies. Hierococcyx varius varius (Vahl, 1797); Hierococcyx varius ciceliae Phillips, 1949.

Description ADULT: Sexes alike, above ashy gray from crown to rump, wing coverts ashy gray, wing gray with a pale patch formed by inner secondaries, tail ashy gray with 4 or 5 blackish bars bordered white distally, a broad subterminal blackish bar and the tip buff; underparts with throat ashy and pale rufous, breast

pale rufous, belly white, belly and flanks with pale rufous wash barred light ashy gray, under tail coverts white, bend of wing white, under wing coverts whitish washed rufous and faintly barred; eye-ring yellow, iris orange to buff to gray, bill yellowish above, greenish below, feet yellow. A distinct subadult plumage is unknown; although collections have a high proportion of birds in plumage as in the following description. JUVENILE: Crown ashy gray and brown with scattered white feathers, nape white, back ashy brown

472 Common Hawk-cuckoo Hierococcyx varius with indistinct rufous bars, tail ashy rufous with blackish bars bordered rufous distally; underparts from throat to upper belly buffy white with broad streaks of brown on breast and broad spots on belly, lower belly and under tail coverts buffy white; bill pale at base and below.

Hierococcyx varius varius (Vahl, 1797); plumage paler; N Pakistan (Rawalpindi), India, Nepal, Assam and Bangladesh; Hierococcyx varius ciceliae Phillips, 1949; plumage darker, more strongly streaked and barred; Sri Lanka.

from 2 to 1.3 kHz with a descending tremolo at the end, in runs of 4 to 6, the phrases in a run becoming higher and louder (Ali 1946, Martens and Eck 1995, P. Alström tape). The song is interpreted in Hindi as “pee-kahan?” or “where is my love?”, and in Marathi, “paos-ala” or “rain’s coming!” (Ali 1996). In Sri Lanka the song appears to be the same (Legge 1880); tape recordings are needed for comparison. In India the song differs from song of Large Hawk-cuckoo H. sparverioides, the first note dropping in pitch, the second note falling (not rising) and the end dropping in pitch (not vibrating around an average pitch) (Figure 6.2). Female gives a strident trilling scream.The cuckoos often sing at night in hot weather, especially on nights with a bright moon.“Brain-fever Bird” is another common name, appropriate for its sound and its insistent calling on hot nights in the Indian plains.

Measurements and weights

Range and status

Northern India (Punjab and Garo Hills) and Nepal: Wing, M (n ⫽ 12) 185–213 (200.5 ⫾ 7.4), F (n ⫽ 7) 187–201 (191.4 ⫾ 5.7); tail, M 160–182 (170.0 ⫾ 7.0), F 146–178 (156.9 ⫾ 10.7); bill, M 21.6–24.8 (23.3 ⫾ 1.2), F 22.1–24.9 (23.4 ⫾ 1.1); tarsus, M 22.5–24.7 (23.8 ⫾ 0.8), F 20.9–24.3 (22.6 ⫾ 1.2) (AMNH, UMMZ). Weight, M (n ⫽ 2) 93–109.6 (101.3), F (n ⫽ 1) 105 (MCZ, USNM). Wing formula, P8 ⬎ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⬎ 10 ⬎ 2 ⬎ 1.

Southern Asia from northern Pakistan (common in West Punjab, uncommon elsewhere), India, Bangladesh and Assam (Garo Hills, Goalpara and Bamanigaon), and Sri Lanka (where locally common in the Dimbula, Dickoya, Dolosbage and Hewata Districts of Central Province and in hill region of Uva Province; perhaps altitudinal migrant moving to lowlands after the breeding season) (Phillips 1948, 1949, Lushington 1949). In India, resident on the plains, locally migratory at higher altitudes and drier habitats, and they move with the monsoon.They are common on the plains, in Punjab Jerdon found them to be the most common cuckoo ( Jerdon 1862), in Maharashtra (Mahabal and Lamba 1987), western Bengal (Inglis et al. 1920) and Kerala (Robinson and

NESTLING: Undescribed. SOURCES: AMNH, BMNH, FMNH, MCZ, RMNH, ROM, UMMZ, USNM,YPM.

Subspecies

Field characters Overall length 33 cm. Uniform gray above, sometimes with pale patch formed by inner secondaries, broad wings, and tail with bands, the gray bands broader than the black bands, throat whitish or gray not black, breast pale rufous and belly indistinctly barred. Adult is pale and lacks the dark streaks on throat and breast of Large Hawk-cuckoo H. sparverioides, juvenile has a pale chin not black as in Large Hawk-cuckoo and is spotted not barred on the belly, and has a pale bill.

Voice Loud, high-pitched shrieking “brain fé-ver”, or “wee-piwhit”, the phrase beginning below 2 kHz, the first note wavering and falling, the second falling

Rufous Hawk-cuckoo Hierococcyx hyperythrus 473 Jackson 1992), and into semi-arid country where they are seasonal in Rajasthan and Gujarat, uncommon elsewhere; not known to migrate to Sri Lanka (Stevens 1925, Phillips 1948, Ali and Ripley 1969, Henry 1971, Kotagama and Fernando 1994, Mukherjee 1995, Grimmett et al. 1999). Reports of breeding in Burma are questioned: Smythies (1940) did not list the bird, but later he did (Smythies 1953) based on a report in the Chin Hills (Hopwood and Mackenzie 1917) of two blue eggs in a nest of Mustached Laughingthrush Garrulax cineraceus (perhaps blue eggs of laughingthrush), and on wartime observations in Arakan (Christison et al. 1946) that did not describe the bird or its song (“brain féver” calls may have been Large Hawk-cuckoo H. sparverioides). Vagrant outside its breeding range, with a specimen record at Masirah I, Oman (Foster 1989, BMNH 1989.5.1), and another reported at Hua Hin, Prachuap Khirikhan Province, in peninsular Thailand (Robinson and Kloss 1921–1924; specimen not traced).

Habitat and general habits Open wooded country, deciduous and semi-evergreen woodland, and pines, mangroves, gardens, orchards, irrigated forest plantations and tea plantations, in semi-arid plains and hill country. In India they are common in deciduous and mixed woodlands from 800 to 1200 m. In Sri Lanka, the dark form ciceliae is resident in the hills above 600 m; paler migrants from India occur in winter in the SW lowlands (Lushington 1949, Inskipp and Inskipp 1985, Martens and Eck 1995, Ali 1996, Grimmett et al. 1999).The cuckoos hunt by perching and watching over the ground, then flutter to the ground to seize an insect, low and swift, rising to perch in a tree, much like an Accipiter hawk (Roberts 1991).

Food Insects, mainly caterpillars and cutworms, nocturnal crickets, grasshoppers, locusts, winged termites, ants, lizards; fruits of wild banyan fig and berries (Ali 1946).

Breeding In India they call in the monsoons, breeding mainly from March to July ( Jerdon 1862, Inglis 1903, Gaston 1981, Ali 1996), with regional variation, near Delhi in March and April, and in Kerala in March (once, young being fed) with most records in October (young being fed, with eggs laid in September), the same period as the peak for host Jungle Babbler Turdoides striatus (Robinson and Jackson 1992, Zacharias and Gaston 2000); in Sri Lanka, January to April (Legge 1880, Phillips 1948). Brood-parasitic. Hosts are babblers, mainly Jungle Babblers in northern and southern India (Davidson 1886, Gaston 1976, Zacharias and Gaston 1983, Gaston and Zacharias 2000) and Sri Lanka (Lushington 1949). Eggs, an oviduct egg was blue, 30.5 ⫻ 20.1 mm (Baker 1906), longer than the blue eggs of Jacobin Cuckoo Clamator jacobinus. Incubation and nestling periods are unknown. The nestling cuckoo evicts the young host from the nest (Ali 1946, 1996, Gaston and Zacharias 2000). Observations of young babblers in a fledged brood with a cuckoo ( Jerdon 1862) suggest that breeding groups pool their broods after they fledge. Fledged cuckoos are fed by babblers for a month after they leave the nest (Gaston 1977, Gaston and Zacharias 2000).The fledged cuckoo’s call “ke-ke” is like that of the babbler (Ali and Whistler 1937).

Rufous Hawk-cuckoo Hierococcyx hyperythrus (Gould, 1856) Cuculus hyperythrus Gould, 1856, Proceedings of the Zoological Society of London, 1856, p. 96. (China, i.e. Shanghai). Other common names: Horsfield’s Hawk-cuckoo (Part), Northern Hawk-cuckoo.

Description ADULT: Sexes alike, above slate gray, wing slate gray, whitish wing patch formed by inner webs of inner flight feathers and coverts, rump slate gray, tail with broad gray bands and narrow dark bands (black, with

474 Rufous Hawk-cuckoo Hierococcyx hyperythrus tan shading distal) and a broad black subterminal band and rufous tip, face and chin slate gray with a white streak from base of bill to throat; underparts with throat white, breast and belly dull rusty red with fine gray streaks, under-tail coverts white; eye-ring yellow, iris yellow to brown, bill black with the base, lower mandible and tip green, feet yellow. SUBADULT: Crown and upper back slate gray, back gray with some rufous feathers, rump slate gray, face and chin slate gray; underparts whitish with light rufous and irregular blackish streaks on breast and flanks, belly and under tail coverts white. JUVENILE: Crown brownish black (not slate gray), nape irregularly blotched white, back indistinctly barred brown and rufous, wing dark brown with rufous bars, whitish wing patch as in adult, rump brown, tail with the broad bands more rufous than gray, tail with rufous tip, face slate gray, chin slate gray, underparts white, the breast and flanks with blackish streaks, eye-ring yellow, iris dark brown, feed yellow. NESTLING: At hatching skin naked, blackish; nestling gape edge and distal palate lemon yellow, the inner mouth pink with dark radial lines (Yoshino 1999). SOURCES: AMNH, BMNH, CM, FMNH, MCZ, MZB, SMF, UMMZ, USNM, ZRC.

Measurements and weights Wing, M (n ⫽ 13) 187–206 (199.0 ⫾ 5.8), F (n ⫽ 5) 193–200 (196.0 ⫾ 2.9); tail, M 127–144 (154, (131.9 ⫾ 5.8), F 126–143 (137.0 ⫾ 8.1); bill, M 19.3–24.2 (21.1 ⫾ 1.5), F 19.1–20.9 (20.3 ⫾ 0.7); tarsus, M 18.2–24 (20.7 ⫾ 1.6), F 18.1–22.5 (20.6 ⫾ 2.0) (AMNH, BMNH, CM, FMNH, MCZ, SMF, UMMZ, USNM). Weight, Japan: M (n ⫽ 1) 115 (“30.7 momme”, a momme being a unit of 3.7 5 g), U (99.0–147.8) (MCZ, Enomoto 1941). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⫽ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 30 cm. Cuckoo with blackish chin, light reddish breast, tail dark with a rufous subterminal band and white tip; wing with a pale patch when seen from above in flight or perched, and often a pale

nape patch.These characters are shared with Philippine Hawk-cuckoo H pectoralis, but H. hyperythrus differs with a dark chin, a more pinkish-red breast with fine streaks, no sharp demarcation of the reddish breast and white belly, and its voice. Differs in the field from H. fugax in its larger size and the extensive reddish area on its breast and belly. Feathered nestlings and juveniles have plumage spotted and streaked not barred, and the barred tail has a rufous subterminal band, unlike that of other young cuckoos in Asia. Young have the mouth yellow, not red as in other cuckoos in their range.The broadly barred tail lacks white spots along the shaft of T1 as occur in Oriental Cuckoo Cuculus optatus. Himalayan Cuckoo C. saturatus, Common Cuckoo C. canorus and Asian Lesser Cuckoo C. poliocephalus.

Voice Song is a two-part buzzy whistle,“zhiu-ichi” or “juichi” or “wee wee-pit”, the first note wavering then held then rising, mostly held for 0.3 sec on one pitch at 3 kHz, the second note dropping slowly then peaking to 4 kHz with a damped oscillation at the end (Figure 6.1). The paired phrasing is like that of Whistling Hawk-cuckoo H. nisicolor and Javan Hawk-cuckoo H. fugax but the notes are harsh and longer.This distinct call of H. hyperythrus was recognized by David and Oustalet (1877) as “Kong-KongYang!” and the bird was then recognized as a distinct species. Other call is a long 3-sec series of short shrill notes rising in pitch then falling at the end, the middle of the series the loudest,“wee taweetaweetaweetaweetaweetaweetaweetaweetatootootootootootoo too”, this crescendo call is similar to that of other hawk-cuckoos H. fugax and H. pectoralis (King 2002, P.Alström recordings). Begging call of fledged young is a repeated buzz (LNS).

Range and status Northeastern Asia, the breeding range in Russia (Ussuria, Sikhote-Alin) (Dement’ev and Gladkov 1966, Neufeldt 1972, Flint et al. 1984, Knystautas and Sibnev 1987), northeastern China (Heilongjiang, from Ussuria and the Yalu River to Beijing) (Cheng 1991) and Japan (southern Hokkaido, Honshu, Shikoku, Kyushu) (Yamashina 1941, Brazil 1981); they are uncommon summer residents. They are passage migrants in Korea (and may breed there),

Philippine Hawk-cuckoo Hierococcyx pectoralis 475

?

from lowlands to 2300 m in Japan. In Russia, they are most common in the southern Maritime region on the middle and upper Iman and the Suptinka River, and the Amur River (Dement’ev and Gladkov 1966, Higuchi 1998).They perch in the crowns of tall trees.

Food Insects, including larvae of hawk moths, sawflies and silkworms (Dement’ev and Gladkov 1966).

Breeding Okinawa and nearby islands, SE China and Vietnam (Delacour and Jabouille 1931, Stresemann 1930b,Yen 1933, La Touche 1934, Ripley 1951, Gore and Won 1971, Tomek 1985, Fiebig 1995, McWhirter et al. 1996,Yoon 2000). In winter they occur in the Philippines (? “hyperythrus” from Catanduanes I, specimen not seen, apparently lost in 1945 war, Dickinson et al. 1991), in the Carolines (Palau Is), in Borneo in the mountains in the north and east (Smythies 1957, 1981, Sheldon et al. 2001), and there are three records in Sulawesi and the Moluccas (Buru) (Siebers 1930, Mayr 1938a,White and Bruce 1986).

Habitat and general habits Taiga forest, deciduous semi-evergreen and evergreen forests in deep ravines and rivers; they occur

In Japan they lay from mid-May until August. Brood-parasitic. Hosts known to have foster-reared the young cuckoos are Japanese Blue Chat (Siberian Blue Robin) Erithacus (Luscinia) cyane, Japanese Robin E. akahigae, Blue-and-white Flycatcher Cyanoptila cyanomelana, Siberian Bluechat (Orange-flanked Bush-robin) Tarsiger cyanurus and Narcissus Flycatcher Ficedula narcissina. Eggs are pale blue, 28 ⫻ 20 mm, 5.6 g (Higuchi 1998, Yoshino 1999). The incubation period is unknown, the nestling period is 19–20 days (Kiyosu 1952). The beginning nestling raises the wings and exposes their lower surface to the foster parent; the bare skin under the wing has the same color and pattern as the open mouth of the begging young (Yoshino 1999).

Philippine Hawk-cuckoo Hierococcyx pectoralis (Cabanis and Heine, 1863) Hiracococcyx pectoralis Cabanis and Heine, 1863, Museum Heineanum, 4(1), 27. (Philippines) Monotypic.

Description ADULT: Sexes alike, above slate gray, wing slate gray, inner secondaries pale gray to buff, tail with broad gray bands and narrow dark bands (black, with tan shading distal) and a broad black subterminal band and rufous tip, face slate gray with a white streak from base of bill to throat; underparts with chin light gray, throat white, breast pale rufous (with or without fine gray streaks), belly and under-tail coverts white, tail light gray with

narrow black and rufous bars and a broad black subterminal band and rufous tip, bend of wing white, under wing coverts unbarred buff white; eye-ring yellow, iris yellow to brown, bill black with the base, lower mandible and tip green, olive below, feet yellow. SUBADULT: (Females lay in this plumage): above dark gray barred rufous, wing dark gray barred rufous, tail barred gray, brown and rufous, more extensively rufous than in adult, the tip black and rufous as in adult, face slate gray; underparts with chin and throat whitish, the breast and belly white streaked black teardrops with a rufous outline.

476 Philippine Hawk-cuckoo Hierococcyx pectoralis JUVENILE: Head blackish with buff edge of feathers, nape white; underparts, chin whitish streaked gray white, the breast and flanks with blackish streaks; iris pale brown. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, CM, CMNH, DMNH, FMNH, FU, JFBM, MCZ, RMNH, ROM, SMF, USNM,YPM.

Measurements and weights Wing, M (n ⫽ 43) 168–189 (174.8 ⫾ 4.7), F (n ⫽ 17) 166–182 (174.2 ⫾ 4.5); tail, M 112–153 (136.0 ⫾ 12.3), F 116–154 (136.7 ⫾ 11.7); bill, M 18.5–23.1 (20.6 ⫾ 1.1), F 19.5–22.8 (20.8 ⫾ 1.0); tarsus, M 17.1–23.3 (20.2 ⫾ 1.1), F 17.6–20.7 (20.1 ⫾ 1.4) (AMNH, ANSP, BMNH, CMNH, DMNH, FMNH, FU, RMNH, ROM, USNM,YPM). Weight, M (n ⫽ 14) 70–80 (76.2), F (n ⫽ 9) 67.3–89.2 (78.6) (CMNH, FMNH, USNM,YPM). Wing formula, P8 ⫽ 7 ⬎ 9 ⫽ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⫽ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 28 cm. Cuckoo with pale chin and unmarked pale throat, unstreaked light rufous breast, dark tail with a rufous subterminal band and white tip; inner webs of inner flight feathers and coverts white barred gray-brown and inner secondaries unbarred pale gray, form a pale patch when seen from above in flight or when perched.These characters are shared with the other hawk-cuckoos Hierococcyx nisicolor, H. fugax and H. hyperythrus except for the underparts.Adult and immature H. pectoralis have a whitish or light gray chin and throat; H. hyperythrus have a black chin that is three times the area of the gray chin in H. pectoralis. In adult H. pectoralis the breast is pinkish rufous and the belly and under tail coverts are white; in adult H. hyperythrus the breast and belly are light rufous. Distinguished in the field by voice, and in the hand by wing length and shape, which is more rounded in H. pectoralis than in H. hyperythrus. Juvenile plumage is streaked not barred, and the tail has a rufous subterminal band, like H. fugax and H. hyperythrus and unlike other young cuckoos in Asia. The broadly barred tail lacks the white spots along the shaft of T1 as occur in other cuckoos including Cuculus saturatus, C. optatus, C. canorus and C. poliocephalus.

Voice Song, a phrase of six short notes, unmodulated whistles that rise then fall, “wheet wheet wheet wheet wheet wheet”, the first three notes each higher and the last three each lower in pitch. The phrases are repeated in a song bout, each phrase in the bout higher in pitch and given with a shorter pause between phrases. The first phrase is at 2.8 kHz and the last phrase is at 4.6 kHz. Late phrases sometimes have more notes and run on and on.After it gives six to eight phrases in a bout, the bird pauses, then it repeats the bout.The notes rise rapidly in pitch and are shorter than the notes of H. fugax and H. nisicolor (Figure 6. 1).The second call is a long crescendo of short buzzy whistles rising then falling through the series, which lasts about four seconds (Scharringa 1999, Kennedy et al. 2000, King 2002, NSA).

Range and status Philippine Islands: Luzon, Negros, Mindanao, Mindoro, Panay, Leyte, Cebu, Sibuyan and Palawan (Siebers 1930, Mayr 1938a, Ripley and Rabor 1958, Rand and Rabor 1960, Gonzales 1983, Kennedy et al. 2000, Miranda et al. 2000). Resident. Uncommon to rare in original dense forest, which is rapidly disappearing in the Philippines.

Habitat and general habits Forest, virgin mossy to dipterocarp forest, on mountain ridges; they occur at higher altitudes to 2400 m (Goodman et al. 1995).

Food Insects, mainly caterpillars; berries (Gonzales 1983).

Javan Hawk-cuckoo Hierococcyx fugax 477

Breeding Females on Luzon and Negros have enlarged ovaries and oviduct eggs in April and May (Ripley

and Rabor 1958, Dickinson et al. 1991, CMNH RSK 5026). Presumed brood-parasitic. The eggs and hosts are unknown.

Javan Hawk-cuckoo Hierococcyx fugax (Horsfield, 1821) Cuculus fugax Horsfield, 1821, Transactions of the Linnean Society of London, 13(1), 178. ( Java) Monotypic. Other common names: Malay Hawk-cuckoo, Hodgson’s Hawk-cuckoo. The name Javan Hawkcuckoo indicates the source of the type specimen. Malay Hawk-cuckoo was used in anglophone faunas when the British were in control of the Malay Peninsula; Hodgson’s Hawk-cuckoo was used for H. nisicolor, and then was taken over for both when the two taxa were combined, but historically the name is not appropriate.

Description

nape irregularly marked white, tail as in adult with a broad black subterminal band and rufous tip, face slate gray, chin slate gray, underparts white, the breast and flanks with blackish streaks, some birds with rufous on sides of neck and breast; eye-ring yellow, iris gray, bill black with yellow base. NESTLING: Undescribed. SOURCES: AMNH, ANSP, BMNH, BPBM, MCZ, MSNG, MZB, RMNH, TISTR, UMMZ, UWBM,WFVZ, ZMA, ZMB, ZRC.

ADULT: Sexes alike, above grayish brown, wing grayish brown, inner secondaries pale gray to buff, tail with broad gray bands and narrow dark bands (black, with tan shading distal) and a broad black subterminal band and rufous tip, tail below light gray with narrow black and rufous bars and a broad black subterminal band and rufous tip, face slate gray with a white streak from base of bill to throat; underparts with chin slate gray, throat white, rest of underparts white with brown streaks and sometimes with trace of rufous on sides of neck and breast, under-tail coverts white, bend of wing white, under wing coverts unbarred buff white; eye-ring yellow, iris brown, bill black with the base, lower mandible and tip green, feet yellow.

Measurements and weights

SUBADULT: Similar to adult but back indistinctly barred rufous, wing blackish with rufous bars, tail with rufous wash in light gray bands, breast and belly whitish with dark brown teardrop-shaped spots, no rufous, feathers with black shaft and white outer barb, tail below more rufous than in adult, with broad rufous (not gray) bands behind the black bands.

Systematic notes

JUVENILE: Crown and back black with feathers narrowly edged buff white, base of feathers white,

Malay Peninsula, Sumatra, Java and Borneo, May to September: Wing, M (n ⫽ 10) 168–180 (174.3 ⫾ 7.2), F (n ⫽ 5) 165–183 (173.4); tail, M 114–148 (135.2⫾ 10.1), F 122–146 (131.2); bill, M 20.0– 25.2 (22.7 ⫾ 1.8), F 20.8–24.8 (22.8); tarsus, M 17.5–23.3 (20.0 ⫾ 2.4), F 17.9–22.3 (20.3) (AMNH, BPBM, MCZ, MSNG, RMNH, UMMZ,WFVZ, ZRC). Weight, Malay Peninsula: U (n ⫽ 2) 74.8–76.5 (75.7) (Wells 1999); Borneo: M (n ⫽ 2) 79.0–86.5 (82.8), F (n ⫽ 1) 74.0 (BPBM, WFVZ, Thompson 1966). Wing formula, P7 ⬎ 8 ⬎ 9 ⫽ 6 ⬎ 5 ⭓4 ⬎ 3 ⫽ 10 ⬎ 2 ⬎ 1.

For a discussion of the species status of hawk-cuckoos in the Hierococcyx fugax complex, see the chapter on song and species.

Field characters Overall length 28–30 cm. Cuckoo with breast streaked, tail dark with a rufous subterminal band and white tip; inner webs of inner flight feathers and coverts white barred gray-brown and the inner

478 Javan Hawk-cuckoo Hierococcyx fugax secondaries unbarred pale gray, form a pale patch when seen from above in flight or when perched, both adult and juveniles with a pale nape patch. Differs from H. nisicolor and H. hyperythrus in brown (not slate-gray) upperparts and in lack of rufous on underparts in all plumages, and “pikwik, pi-kwik” song is lower-pitched than in H. nisicolor with the second note about as loud as the first, not louder as in H. nisicolor. In the hand, differs from H. nisicolor in the more rounded wing and the longer bill. Chasen and Kloss (1927) reported the under wing coverts unmarked or sparsely marked in H. fugax; and marked with transverse bands in H. nisicolor; in fact the coverts are unmarked in all age groups in both forms (AMNH, UWBM). A pale wing patch also occurs in the less boldly-marked Common Hawk-cuckoo H. varius. Feathered nestlings and juveniles have plumage streaked not barred, and barred tail has a rufous subterminal band, unlike in other young cuckoos in mainland Asia except other hawkcuckoos. The broadly barred tail lacks white spots along the shaft of T1 as occur in Cuculus saturatus, C. optatus, C. canorus and C. poliocephalus.

Voice Song is a series of shrill slurred doublets (or triplets) “pi-kwik”, each note rising in pitch, the first note either higher or lower in pitch than the second, the notes about 3–4 kHz (Figure 6.1).The first note rises gradually in pitch with a peak energy between 3 and 4 kHz, then jumps to a high ending; the second note is similar. Songs are either repeated at the same pitch or rise through a bout, the final songs at a peak amplitude as high as 4.2 kHz for the first note (Pasoh Forest Reserve, Malay Peninsula, Sharringa 1999). Other call is a long, rapid sequence of shrill buzzy whistles “pee” that rise up the scale in crescendo then drop at the end in the Malay Peninsula and Borneo, “trrrrrtitititititirrrtrrr”, the crescendo phase with some higher long screeches with the short “pee” notes. The songs are similar in Thailand (Ban Khao), the Malay Peninsula, Sumatra and Borneo (Scharringa 1999, Wells 1999, King 2002, P. Alström recordings, P. Round recordings, LNS, NSA).

Range and status Southern Burma (Tenasserim south of Tavoy), Thailand, the Malay Peninsula, Sumatra, the Mentawai Islands, Rhio Archipelago, Bangka; western Java, Batu Islands, Billiton, Karimata Islands and Borneo (Siebers 1930, Kuroda 1936, Junge 1948, Smythies 1940, 1953, 1957, 1981, van Marle and Voous 1988, Wells 1999, Kemp 2000). Resident in most of the range. Some migration or dispersal occurs in the Malay Peninsula where movement is indicated by a grown juvenile (UWBM 67529) found in July 2000 in Singapore, a well-birded area where this species does not breed (Hails 1987), and an adult female came to a light at night in Borneo (Sheldon et al. 2001, WFVZ). In Sabah, several kinds of hawk-cuckoos occur in winter on Mt Kinabalu, including H. hyperythrus and H. nisicolor which winter there and H. fugax which is a breeding resident. Uncommon through the range. Some old records of H. fugax may refer to H. vagans, as the two species were not differentiated until 1877.

Habitat and general habits Forests. In the Malay Peninsula they occur mainly in lowland forest, in Sumatra from lowland forests to 1400 m (van Marle and Voous 1988), in Borneo they are in woodland and scrub, lowlands and hills, the resident form H. fugax in lowland forests and cocoa plantations and as high as 1600 m in the Kelabit uplands, and wintering hawk-cuckoos in the mountains, H. hyperythrus in the north and east, H. nisicolor

?

Whistling Hawk-cuckoo Hierococcyx nisicolor 479 on Mt Kinabalu, in Kelabit uplands to 2800 m, and in SE Borneo (Smythies 1981, Sheldon et al. 2001).

Food Insects, mainly caterpillars, locusts, cicadas, beetles, butterflies; and fruits and berries (Smythies 1981, Sody 1989).

Breeding In peninsular Thailand, fledglings called, one with a “burr” and the other with a harsh “yelp” in July; on Penang a fledgling was fed by its host in August, and in the Malay Peninsula, adults call from February through August and around the clock during moonlit weeks in April (Wells 1999). In Sumatra a growing fledgling was taken in June ( Junge 1948),

and in Borneo a female laid in April (Sharpe 1890) and an attended fledgling was seen in August (Cranbrook and Wells 1981). Brood-parasitic. In Malay Peninsula a fledgling was fed by White-rumped Shama Copsychus malabaricus (Wells 1999), in Borneo a female laid in nest of Grey-headed Canary-flycatcher Culicicapa ceylonensis (Whitehead, in Sharpe 1890) and a fledgling (this cuckoo species?) was attended by Black-throated Babbler Stachyris nigricollis (Cranbrook and Wells 1981). Eggs, in Borneo one attributed to this cuckoo was creamy white with brown and gray spots in a wreath at the large end, 22 ⫻ 16 mm (Whitehead, in Sharpe 1890, Schönwetter 1964). The incubation and nestling periods are unknown.

Whistling Hawk-cuckoo Hierococcyx nisicolor (Blyth, 1843) Cuculus fugax nisicolor Blyth, 1843, Journal of the Asiatic Society of Bengal, 12, p. 943. (Nepal) Monotypic. Other common names: Hodgson’s Hawk-cuckoo (Oates and Blanford 1895), Himalayan Hawkcuckoo (Baker 1927).The name Whistling Cuckoo was used by Blyth (1842);Whistling Hawk-cuckoo describes the song of this bird better than the song of other hawk-cuckoos.

Description ADULT: Sexes alike, above gray brown, wing slate gray, inner secondaries pale gray to buff, tail with broad gray bands and narrow dark bands (black, with tan shading distal) and a broad black subterminal band and rufous tip, tail below light gray with narrow black and rufous bars and a broad black subterminal band and rufous tip, face slate gray with a white streak from base of bill to throat; underparts with chin slate gray, throat white, breast light rufous thinly streaked with rufous and brown on sides, lower breast and upper belly light rufous, lower belly and under-tail coverts white, bend of wing white, under wing

coverts unbarred buff white; eye-ring yellow, iris brown, bill black with the base, lower mandible and tip green, feet yellow. SUBADULT: Similar to adult but back indistinctly barred rufous, wing blackish with rufous bars, tail with rufous wash in light gray bands, breast and belly whitish with dark brown teardropshaped spots and streaks with rufous edge, feathers with black shaft, rufous inner barbs and white outer barbs, tail below more rufous than in adult, with broad rufous (not gray) bands behind the black bands. JUVENILE: Crown and back black with feathers narrowly edged buff white, base of feathers white, nape irregularly marked white, tail as in adult with a broad black subterminal band and rufous tip, face slate gray, chin slate gray, underparts white, the breast and flanks with blackish streaks, some birds with rufous on sides of neck and breast; eye-ring yellow, iris gray (some with a brown iris ring), bill black with a yellow base. NESTLING: Undescribed.

480 Whistling Hawk-cuckoo Hierococcyx nisicolor SOURCES: AMNH, ANSP, BMNH, MCZ, RMNH, UMMZ, USNM, UWBM,YPM, ZMA, ZMB.

Measurements and weights India, Nepal and Burma:Wing, M (n ⫽ 6) 174–185 (178.3 ⫾ 4.6), F (n ⫽ 5) 176–184 (181.0 ⫾ 5.0); tail, M 122–143 (135.3 ⫾ 9.6), F 128–143 (134.8 ⫾ 7.4); bill, M 20.4–20.8 (20.6 ⫾ 0.2), F 20–22.3 (20.9 ⫾ 1.0); tarsus, M 18.9–20.1 (19.0), F 18.1– 22.2 (19.4) (AMNH, BMNH, UMMZ, USNM). Weight, Malay Peninsula, migrant: U 69.1–93.0 (most ⬍ 80.0) (Wells 1999)). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⫽ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 28–30 cm. Cuckoo with slate-gray upperparts, breast streaked, tail dark with a rufous subterminal band and white tip; inner webs of inner flight feathers and coverts white barred gray-brown and the inner secondaries unbarred pale gray, form a pale patch when seen from above in flight or when perched, both adult and juveniles with a pale nape patch.These characters are shared with Rufous Hawk-cuckoo H. hyperythrus which is more extensively rufous below. Differs from Javan Hawkcuckoo H. fugax in the presence of rufous underparts in adult and subadult, and in the hand in the shorter bill and more pointed wing. A pale wing patch also occurs in the less boldly marked Common Hawkcuckoo H. varius. Feathered nestlings and juveniles have plumage streaked not barred, and the barred tail has a rufous subterminal band, unlike in other young cuckoos in mainland Asia except for H. hyperythrus and H. fugax.The broadly barred tail lacks white spots along the shaft of T1 as occur in Oriental Cuckoo Cuculus optatus, C. saturatus, C. canorus and C. poliocephalus.

Voice Song is a shrill thin piping whistle “gee-whiz”, each note rising in pitch, the first note lower in pitch than the second, the “gee” about 4–5 kHz, c. 8 notes in 10 sec and louder than the first, the phrase repeated up to 20 times, sharp and crisp, not modulated or buzzy (Figure 6.1). The first note is held

most of its length on a single pitch (between 4.6 and 5.4 kHz) before the note rises to a peak, the second note is shorter and lower in pitch (with peak energy between 4.4 and 4.6 kHz), rises through its length and is not held on one pitch.The other call is a long, rapid sequence of shrill buzzy whistles “pee” that rise up the scale in crescendo then drop at the end, “trrrrr-titititititirrrtrrr”, the crescendo phase with some higher long screeches with short “pee” notes. A churred call perhaps an alarm is given in response to playback of song (P. Round). The songs are similar in southern China (Sichuan), northern India, Nepal, Bhutan and Thailand (King 2002, B. King recordings, P. Alström recordings, P. Holt recordings, P. Round recordings, LNS, NSA).

Range and status China south of lower Yangtse River, in the upper Yangtse basin south of 30° N in Sichuan, Yunnan, Guangxi, Guangdong and Hainan (Stresemann 1930b,Yen 1933, Cheng 1991, Li 1991), and in the eastern Himalayas in India from Darjeeling to Nepal, Bhutan and Assam (Tymstra et al. 1997, Bishop 1999, Grimmett et al. 1999, J. Cox pers. comm.), Burma (Chin hills, Shan hills, Karen hills: Smythies 1940, 1986) and Thailand from Nan Province south to Khao Yai (Deignan 1945, Lekagul and Round 1991, P. Round recordings), Laos and Vietnam (Tonkin and Annam) (Thewlis et al. 1996, Robson 2000a). Migrant in most of range. Birds in Sikkim were taken from April through October (AMNH, BMNH). Birds in southern Burma may be passage migrants; birds in Laos, Tonkin and

?

Black Cuckoo Cuculus clamosus 481 Annam may be residents. During migration season birds move through the Malay Peninsula and come to lights at night, striking a lighthouse in the Malacca Straits, and in winter birds occur in the Malay Peninsula (Gibson-Hill 1949b, Medway and Wells 1976, Wells 1999, UWBM) and in Sumatra, Bangka, Belitung, western Java, and Borneo (Kuroda 1936, Deignan 1945, Smythies 1957, 1981, Mees 1986, van Marle and Voous 1988).

and H. nisicolor on Mt Kinabalu, in the Kelabit uplands to 2800 m, and in SE Borneo (Smythies 1981, Sheldon et al. 2001).

Food Insects, mainly caterpillars, also locusts, cicadas and beetles (Baker 1927).

Breeding Habitat and general habits Deciduous semi-evergreen and evergreen forests, secondary forest, damp ravines with pines, bamboo thickets and plantations, active in bushes and understory (Smythies 1940). Breeding elevation ranges from 600 to 1800 m in Assam (Baker 1927), 1000 to 2000 m in Bhutan (Nobji and Gasa valleys) (Tymstra et al. 1997), in hills to 1000 m in Burma, and below 350 m in northern Thailand. In winter, in Borneo three closely-related hawk-cuckoo species live in woodland and scrub, lowlands and hills, the resident form H. fugax in lowland forests and cocoa plantations and as high as 1600 m in Kelabit uplands, and wintering hawk-cuckoos in the mountains, H. hyperythrus in the north and east,

In India birds call from May to August (Baker 1937) and lay from June (Sikkim, oviduct egg taken 5 June by Mandelli, in Hume and Oates 1890) to July, an egg was found as late as September (Baker 1906, 1934). In northern Thailand a juvenile was taken in July (Deignan 1945). Brood-parasitic. In India, the hosts are Nepal Shortwing Brachypteryx leucophrys, Small Niltava Muscicapa macgrigoriae and other flycatchers and babblers (Hume 1873, Baker 1934, Becking 1981). Eggs in India and Sikkim are uniform olive brown to green, darker at the thick end, 22.5 ⫻ 15.4 mm (Baker 1934), 24 ⫻ 16 mm (Ali and Ripley 1969). The incubation and nestling periods are unknown.

Genus Cuculus Linnaeus, 1758 Cuculus Linnaeus, 1758, Systema Naturae (ed. 10) 1, 110.Type, by tautonymy, Cuculus canorus Linnaeus. L. Cuculus, the cuckoo. Small to medium-sized cuckoos with whistled songs, and round nostril. Type, by tautonymy, Cuculus canorus Linnaeus, 1758. Most species are gray above and barred below, the barring absent in some forms of African

Black Cuckoo C. clamosus. Widespread in Old World from Africa, Eurasia to the Philippines and Indonesia; nonbreeding migrants occur to New Guinea and Australia. Most Cuculus species have a pointed wing, barred underparts, unstreaked juvenile plumage and a whistled song. Nine species.

Black Cuckoo Cuculus clamosus Latham, 1801 Cuculus clamosus Latham, 1801, Index Ornithologicus, Suppl., p. xxx. (Cape of Good Hope ⫽ Cradock) Polytypic.Two subspecies. Cuculus clamosus clamosus Latham, 1801; Cuculus clamosus gabonensis Lafresnaye, 1853.

Description ADULT: Sexes alike or similar, black, above glossed greenish, wing dull slate, tail black narrowly tipped white, outer tail feathers sometimes with white or buff spots or bars; underparts either black from

482 Black Cuckoo Cuculus clamosus throat to under tail coverts (C. c. clamosus) or chin to breast rufous and belly with white bars (C. c. gabonensis), bend of wing black with white flecks, under wing coverts slate, inner vane of primaries barred white; eye-ring black, iris brown, bill black, feet black; males usually unbarred and females often barred rufous to light brown on breast, belly and (especially) under tail coverts. In southern and East African C. c. clamosus, adult plumage is barred in 4 of 23 males and 7 of 12 females (UMMZ, ZMUC). JUVENILE: Above black, tail with black tip and the tips of the rectrices more pointed than in the adult, wing dull slate, underparts black. Birds molt directly from black juvenile plumage into adult plumage with the tail tipped white and (in C. c. gabonensis) a rufous breast and underparts with black and white. NESTLING: At hatching the nestling is naked with pale brownish pink skin which darkens to purplish black; the inside of the mouth is pink. SOURCES: AMNH, ANSP, BMNH, BWYO, CM, CU, FMNH, MVZ, UMMZ, USNM, ZMUC, ZFMK.

Subspecies Cuculus clamosus clamosus Latham, 1801; sexes alike, underparts black, sometimes barred blackish and rufous or buff (the light bars narrower than the black bars); Ethiopia, E and C Kenya, Tanzania, southern Zaire, Angola south to NE Namibia, Zimbabwe, and eastern South Africa; Cuculus clamosus gabonensis Lafresnaye, 1853; chin to breast rufous, remainder of underparts barred black and white (sometimes belly white without the bars, or with the bars incomplete), under-tail coverts usually barred, female duller;West and Central Africa from Liberia, Ghana, Nigeria and southern Sudan through Zaire, Uganda and western Kenya. Typical gabonensis have black bars of the belly narrower than the white bars. Females often have fine black bars on the rufous throat and breast; males lack the black bars. Some adult gabonensis have the greenish eye-ring and feet yellowish to pale flesh (AMNH). Birds in the Central African

forests are variable in plumage, and some have been named as subspecies, but the regional distributions of these color phases overlap; within these areas cuckoos are variable in plumage (noted also by Friedmann 1930, 1948, 1966) and they are better known as plumage phases. The phase “jacksoni” Sharpe 1902 has black bars wider than the white to rufous-buff bars, and throat and breast plumage a mix of rufous and black. Birds intermediate in plumage between barred “jacksoni” and black C. c. clamosus with rufous and black bars from breast to belly occur from Sierra Leone (MCZ, June 1952; BMNH, May 1903) and Fouta Djalon region of Guinea (UMMZ photo, Oct 1999) to Nigeria and east to Eritrea, Ethiopia, Uganda and western Nigeria (BMNH, CU, MSNG, USNM). Birds in the “mabirae” van Someren 1915 plumage have lower breast and belly all white or white sparsely marked with black a rufous throat and breast and a white belly, or a white belly sparsely marked with black bars or with black bars only on the upper flanks (not “intergrading” between gabonensis and solitarius as in Sclater and Mackworth-Praed 1919). These occur in forested east, central and western Zaire, Gabon and Cameroon (Efulan, Kribi, Nko’olong) (FMNH,AMNH); the first one collected and named was from Mabira forest in Uganda but none have been seen there recently.The variable plumages were once thought to indicate different ages (Bannerman 1921). However, the occurrence of birds with a mix of black plumage (the juvenile) and the other forms shows a single molt from juvenile to adult plumage, as described for gabonensis by Bates (1911). A distinctive sub-adult plumage is known among cuckoos only in the Asian hawk-cuckoos, Hierococcyx. Birds in all parts of their range have the same song. In addition, the juveniles in all populations are uniformly black without a white tip on the tail. Birds in all parts of their range overlap considerably in size as well. Within southern Africa, where adult birds are mostly or entirely black, there is no evidence of assortative mating of the all-black birds or of the partlybrown-barred birds, and these plumages are more closely associated with sex than with geographic locality. In forested regions of central Africa where some birds are nearly all white below (“mabirae” ) and

Black Cuckoo Cuculus clamosus 483 others are barred black and white (gabonensis), the plumages do not appear strongly associated with sex (though there are few sexed female specimens in collections), in Uganda both “gabonensis” and “jacksoni” birds with intermediates occur, and in Kenya where both barred and unbarred black birds occur, many adult plumages are intermediate between the “jacksoni” forest birds in the gabonensis complex and the unbarred black clamosus birds of open woodlands.

Measurements and weights C. c. clamosus, South Africa: Wing, M (n ⫽ 10) 163–180 (174.3 ⫾ 4.7), F (n ⫽ 6) 173–186 (177.5 ⫾ 4.4); tail, M 152–166 (157.2 ⫾ 5.5), F 142–162 (152.5 ⫾ 8.5); bill, M 21.8–24.2 (23.0 ⫾ 0.7), F 21.1–23.5 (22.5 ⫾ 0.8); tarsus, M 18.3–22.4 (20.0 ⫾ 1.3), F 18.2–23.1 (20.9 ⫾ 2.1) (UMMZ); C. c. gabonensis, Zaire (Uele, Ituri, Kivu, Kasai): Wing, M (n ⫽ 14) 161–187 (167.9 ⫾ 4.4), F (n ⫽ 4) 156–166 (160.8 ⫾ 4.6); tail, M 127–162 (140.1 ⫾ 8.1), F 140–162 (147.5 ⫾ 9.9); bill, M 19.8–3.4 (21.6 ⫾ 1.0), F 20.4–24 (21.8 ⫾ 1.6); tarsus, M 17.1–21.2 (19.9 ⫾ 1.2), F 18–19.4 (18.7 ⫾ 0.6) (AMNH). Weight, South Africa, Botswana, Zambia: M (n ⫽ 10) 78–103.5 (90.2), F (n ⫽ 2) 86–88.3 (after laying an egg) (87.2), F, egg in oviduct (n ⫽ 1) 102.7 (UMMZ); Zaire, U (n ⫽ 3) 75–94 (87.3) (Louette and Herroelen 1993); Nigeria, M (n ⫽ 5) 80.9–88.0 (84.5) (CU). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⫽ 1 0 ⬎ 2 ⬎ 1.

Field characters Overall length 31 cm.A large cuckoo with all-black plumage and a distinctive whistled song. In flight seen from below, bird sometimes has a white wing bar, the white inner vane of the flight feathers. Birds in West and Central Africa have a rufous breast and a whitish belly heavily to sparsely barred with black, and differ from Red-chested Cuckoo C. solitarius in rufous chin and throat (throat gray in C. solitarius). Both the open woodland form and the forest form of C. clamosus have a slow, rising threenote whistle (descending notes in C. solitarius).The juvenile plumage in both subspecies is uniformly black, without pale edges on feathers of the upper-

parts and without bars below (pale-edged above, and the belly and under tail coverts are barred black and white in juvenile C. solitarius).

Voice A loud, mournful slow whistle “whoo whoo whee” or “pa-pa’s boy”, the second note shorter and higher, the third rising in pitch a third of an octave. Call in aggressive context is a long call, a harsh crescendo “gagagaGAGAGA!” rising in pitch to a climax then dropping in pitch and volume, a ghoulish cry. Bates (1930) called this the “hurry-hurry” call. Female call, a “kwik-kwik-kwik-kwik!”. Songs are the same throughout the species range and across the plumage forms in forest and open woodland ( Jackson 1938, Chapin 1939, Marchant 1942, Bannerman 1951, Keith and Gunn 1971, Chappuis 1974, 2000, Stjernstedt 1993). The begging call of the young cuckoo is a thin, wispy “sweeet” (Skead 1946).

Range and status Africa throughout sub-saharan region. C. c. clamosus are migratory in southern Africa, where they sing mainly September to December (Niven and Niven 1966, Benson and Benson 1977, Rowan 1983, Tarboton et al. 1987,Vernon et al. 1997), in Zambia mainly October to March (Aspinwall and Beel 1998), in all months locally in East Africa (Zimmerman et al. 1996).They are generally absent in the dry parts of East Africa (N Kenya,W Tanzania) (Stevenson and Fanshawe 2002). Winter in Africa within 12° north to south of the equator from S Ethiopia, Sudan and Central African Republic to Sierra Leone and Guinea, mainly north of the forest zone south to Enugu, Ibadan and Cape Coast. Adult cuckoos in black plumage occur in Central and West Africa from January to September, when they are not breeding in southern and East Africa, and there are no breeding records in West Africa (Friedmann 1978, Thiollay 1985, Grimes 1987, Irwin 1988, Elgood et al. 1994, Cheke and Walsh 1996, Gatter 1997, Claffey 1998; BMNH). On the other hand, there are no breeding records of the rufous-breasted forest cuckoos west of Cameroon, because of the scarcity of field observations in forests in West Africa; a black-plumaged bird (CU) taken near Ibadan, Nigeria, had large testes,

484 Black Cuckoo Cuculus clamosus

Food Insectivorous, mainly caterpillars (hairy and smooth; in the Eastern Cape especially larvae of the barred eggarlet Bombycomorpha bifasciata); also takes termites, grasshoppers, beetles, grubs, winged ants; ticks, birds’ eggs, small birds (nestlings?) (Chapin 1939, Friedmann 1948, Skead 1951, 1995, Brooke 1965b, Rowan 1983, AMNH).

Displays and breeding behavior

and birds in black plumage may breed in savannas in West Africa. C. c. gabonensis are resident in forests in West and Central Africa; at Mt Kupé, Cameroon, they occur in all months. Common in open and closed forests in the lower Guinea region of West Africa and equatorial Central and East Africa and in open woodlands in southern Africa. An adult ringed at Amanzi, Eastern Cape, in November 1963 was recovered there as a breeding male in November 1965. It had a black breast and belly with buff barring on under tail coverts (UMMZ 216712), so barred plumage does not necessarily indicate a young bird.

Habitat and general habits Forest, open woodland, miombo, riparian woodland, acacia thicket; they occur in open woodlands and avoid evergreen forest and thicket in southern Africa, except along the coastal belt in Eastern Cape where they are mainly in thickets and valley bushveld. They occur in plantations and trees around human habitations. In contrast to their habitat in southern Africa, they are more a forest bird than Red-chested Cuckoo C. solitarius in Central and West Africa, although the two cuckoos live together in some areas such as Tai NP, Ivory Coast, and forest localities in Uganda (Bates 1930, Friedmann 1948, Urban and Brown 1971, Irwin 1988, Gartshore 1989, Lewis and Pomeroy 1989, Gatter 1997, Rossouw and Sacchi 1998,Vernon et al. 1997, Struhsaker 1998). In feeding they search through foliage in trees, hawk flying insects from the air, and take caterpillars on the ground (Skead 1995).

Male sings in territory, and is attracted to playback of song or to a human whistle.A male sometimes sings in a tree where one or two other males perch quietly, all of them with enlarged testes (UMMZ).Two males attended to an ovulating female (RBP). The female removes an egg from the host nest when she lays her own (Friedmann 1948). Courtship feeding has been observed (Bannerman 1951).

Breeding In southern Africa they lay October to December (Skead 1951, 1995, Benson and Benson 1977, Irwin 1981, Maclean 1993, Dean 2000), in northern Zambia males sing in miombo woodland when trees are in fresh foliage before the rains and these birds respond by approaching a whistled imitation of song (3 Sept 1966, RBP); in Tanzania they breed in March and April and in Kenya and Uganda the season is not precise across the region but varies with the more local rainfall (Friedmann 1948, Brown and Britton 1980). There are few breeding records in Central Africa (Bates 1927, Chapin 1939, Serle 1965); in Central African Republic a female C. c. gabonensis had a large ovary in June (Friedmann 1978). In Cameroon, at Mt Kupé they sing April to November (Bowen 2001). A female lays on alternate days, usually in a series of 4–5 eggs, in a good year as many as 22 eggs in a season (Payne 1973a, 1974). In South Africa, adult cuckoos begin the postbreeding molt at the end of the breeding season while still on the breeding grounds (Payne 1969a). Brood-parasitic. Hosts, C. c. clamosus use mainly boubou shrikes Laniarius (Tropical Boubou L. aethiopicus, Southern Boubou L. ferrugineus, Gabon Boubou L. bicolor and Crimson-breasted Shrike L. atrococcineus) in scrub habitats in southern Africa (Friedmann

Red-chested Cuckoo Cuculus solitarius 485 1967, Payne and Payne 1967, Jensen and Jensen 1969, Jensen and Clinning 1974, Rowan 1983, Irwin 1988, Pryce 1989, Skead 1995, Skinner 1996). In southern Cameroon a forest cuckoo C. c. gabonensis fledgling was brought to Bates (1911) by a man who called it a young Sooty Boubou Laniarius leucorhynchus and saw it in company with the boubou. The collector returned with the boubou, which had been “still crying for its child”, and the same kind of food was found in the stomachs of the boubou and the young cuckoo. A fledged cuckoo that associated with Cassin’s Malimbe Malimbus cassini was perhaps adopted after it had fledged (Dowsett-Lemaire 1996). The eggs of C. c. clamosus are whitish or pale green with reddish brown marks, similar to eggs of boubous (Tarboton 2001). In females with a hard, pigmented egg in the oviduct, 3 eggs were light blue with brown marks. Two of

these eggs were 26.8 ⫻ 18.6 mm (UMMZ 216709 at Kei Road, Eastern Cape), and 25.4 ⫻ 19.1 mm (UMMZ 216718 at Marble Hall,Transvaal, a female that was chased by a Crimson-breasted Shrike and laid the egg when she was captured, the fresh egg was 5.0 g). Eggs from laying forest cuckoos C. c. gabonensis in Cameroon were the same, creamy white with spots and speckles of light gray and brown or lilac, one was 23.5 ⫻ 17 mm (Bates 1909, 1927, 1930). The incubation period is 14 days. The nestling evicts the host eggs two days after it hatches. By day 4 the sheaths of the main feather tracts emerge, by day 6 the nestling is in pinfeathers, by day 9 the eyes are open, and by day 16 the nestling is well feathered. The nestling period is 20–21 days, and the fledged young cuckoo remains with its foster parents for 19–43 days (Skead 1951, Jensen and Clinning 1974, Rowan 1983).

Red-chested Cuckoo Cuculus solitarius Stephens, 1815 Cuculus solitarius Stephens, 1815, in Shaw’s General Zoology, 9, 83. (Caffraria and Madagascar ⫽ Eastern Cape Province) Monotypic.

Description ADULT: Sexes similar, male, above dark gray, wing slate gray, tail blackish tipped white, rectrices black in center with white spots along the shaft, T1 and T2 dark along the edge,T3 to T4 with white spots along the edge,T5 with white bars; underparts, chin gray, throat unbarred gray, breast rufous, lower breast and belly buffy white barred black, the black bars half the width of the white bars (1.6 vs 3.4 mm), under tail coverts buffy white either unbarred or with a few incomplete black bars, the tail appearing barred when seen from below, bend of wing white with black bars, under wing coverts whitish barred black, inner vane of primaries barred white; eye-ring yellow green, iris brown, bill black, feet yellow. Female, upper breast sometimes barred, breast paler rufous than in male, some females with hardly any rufous. JUVENILE: Above, crown black with fine buffwhite tips, nape with a few white feathers, back

blackish with white margins of feathers, tail black with broad white spots and tip,T5 with white bars, tips of rectrices more pointed than in the adult; underparts, throat to upper breast black with white margins of feathers, belly barred black and white (the black bars narrower than the white bars, 2.6 vs 4.0 mm); eye black, bill black, feet orange. NESTLING: Hatchling is naked, skin black to dark brown; inside of mouth is yellow and darkens to orange, and the feet are yellow. SOURCES: AMNH, BMNH, BWYO, CM, FMNH, MCZ, MVZ, RMNH, UMMZ, USNM, ZMUC.

History and subspecies Stresemann (1924a) considered this cuckoo to be a color phase of Black Cuckoo C. clamosus. Nevertheless, they do not interbreed, their songs and juvenile plumages differ, and they live in different habitats (C. solitarius in forests and C. clamosus in open woodland in southern Africa, the reverse in East, Central and West Africa). Amadon (1953) described birds of the Gulf of Guinea islands as a long-billed subspecies

486 Red-chested Cuckoo Cuculus solitarius C. s. magnirostris; but long-billed birds also appear on the mainland (Eisentraut 1973). Bill lengths of males on Bioko (n ⫽ 5) range from 19.1–23.0, and males in Ituri and the upper Congo River (n ⫽ 12) from 19.0–21.4 (AMNH). These populations are too far apart to support the idea that longbilled birds on the African mainland are migrants from the Gulf of Guinea (Dowsett 1993).

Measurements and weights Wing, M (n ⫽ 20) 164–180 (172.0 ⫾ 5.6), F (n ⫽ 11) 156–187 (169.0 ⫾ 10.0); tail, M 130–167 (144.7 ⫾ 10.4), F 130–154 (142.8 ⫾ 9.1); bill, M 20.2–22.7 (20.7 ⫾ 1.2), F 17.6–21.6 (20.2 ⫾ 1.2); tarsus, M 17.7–24 (19.9 ⫾ 1.6), F 17.6–21.6 (19.0 ⫾ 1.1) (AMNH, UMMZ). Weight, M (n ⫽ 4) 67–82.6 (77.5), F (n ⫽ 1) 70.4 (UMMZ). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 3 ⫽ 1 0 ⬎ 2 ⬎ 1.

Field characters Overall length 31 cm. Large cuckoo with dark upperparts, barred underparts, and a rufous breast (pale and barred in some females). Tail differs from Common Cuckoo C. canorus and African Cuckoo C. gularis by T1 to T5 with black not gray edge, and by white on central T1 limited to spots along the shaft. Chin and throat gray (rufous in rufous-breasted or rufous-barred forest forms of Black Cuckoo C. clamosus gabonensis). Juvenile differs from that of C. clamosus in the black plumage with feathers narrowly edged and tipped white, the tail with white spots and bands, and the belly and under tail coverts barred black and white (all black in C. clamosus).

Voice Song, loud, emphatic, three notes dropping in pitch, “Piet-my-vrow!” or “IT-will-rain!”; the call “Pit me frow” was described by Levaillant (1790) as a male calling for his mate, which Levaillant’s colleague Pit had just collected. Often sings at night. Female call a loud “kwik-kwik-kwik-kwik!” (North 1958, Chappuis 1974, 2000, Stjernstedt 1993, RBP). Begging call by young birds is a highpitched insect-like buzz, slightly decreasing in pitch (Salewski and Grafe 1999).

Range and status Africa from Senegal, Nigeria, southern Sudan and Ethiopia, south through Zaire,Angola, Zambia, Zimbabwe, Mozambique, and eastern and southern South Africa along the coast to the Cape. Intra-African migrant, migratory in southern Africa.They occur on the coast of East Africa between October and April (Britton 1980a,b), while in southern Zaire, Zambia, Zimbabwe, Malawi and South Africa they arrive in September and leave in March (Verheyen 1953, Benson and Benson 1977, Irwin 1981,Vernon et al. 1997, Aspinwall and Beel 1998). In West Africa perhaps migratory, in Nigeria they move from coastal areas northwards to northern savannas during the northern rains (Elgood et al. 1994). On Gulf of Guinea islands, they are resident (Amadon 1953, Eisentraut 1973, Pérez de Val 1996, Pérez de Val et al. 1997). Common.

Habitat and general habits Forest, open woodland, riparian woodland, thorn scrub, and in montane areas to 3000 m. A forest bird in west, central, and southern Africa, they are more an open woodland bird in East Africa (Friedmann 1930, 1948, Friedmann and Loveridge 1937,Thiollay 1985, Irwin 1988, Gatter 1997). In East Africa they are mainly above 1000 m and within 500⫹ mm rainfall areas, and they breed in semi-arid acacia savanna (Lewis and Pomeroy 1989, Zimmerman et al. 1996).

Food Insects, mainly hairy caterpillars, and beetles, grasshoppers, flying termites; spiders, centipedes, millipedes, snails, slugs, small frogs and lizards; berries,

Red-chested Cuckoo Cuculus solitarius 487 and eggs of its hosts (Chapin 1939, Rowan 1983, Irwin 1988; UMMZ 216703 from Brits, Transvaal, bird not laying, throat plumage stained with yolk).

Displays and breeding behavior Male sings in territory and is attracted to playback of song. Male also sings in tree where one or two other males perch quietly. Copulation occurs without obvious display, male perched near a female mounted her, the pair separated, the male flew away then returned with a caterpillar which he presented to female; she accepted and the pair mated again in a bout of courtship feeding. A female repeatedly approached a nest defended by a host, and when successful perched near the nest, flopped onto the nest with wings spread, and after five seconds flew, having removed a host egg and laid its own egg (Symons 1978). A female cuckoo lays about 20 eggs in a breeding season (Friedmann 1948, Brooke 1964, Payne 1973a, 1974).

Breeding and life cycle During early rains; the birds call in southern Senegal and The Gambia from June to August, a female had an egg ready to lay in July, and a juvenile was seen in October (Morel and Morel 1990, Barlow et al. 1997); in inland Liberia they call from December to April (Colston and Curry-Lindahl 1986), in eastern Liberia they call from March to May and September to November (Gatter 1997) and in Benin,Togo and Nigeria from early April to September (Elgood et al. 1994, Cheke and Walsh 1996, Claffey 1998). In Cameroon a female had a developed egg in August (Serle 1950b). In East Africa the breeding season varies with the rains, in the wet months March to May ( Jackson 1938, Skinner 1978, Brown and Britton 1980), in Ethiopia perhaps April-July (Urban and Brown 1971), in southern Africa they breed from October to January (Benson et al. 1971, Dean 1971, Benson and Benson 1977, Irwin 1981, 1988, Lewis and Pomeroy 1989, Maclean 1993). Brood-parasitic. Hosts, mainly thrushes, robinchats and alethes (Friedmann 1948, 1967, van Someren 1956, Payne and Payne 1967, Oatley 1970, Rowan 1983, Nikolaus 1987, Irwin 1988, DowsettLemaire 1996, Tarboton 2001). In East Africa the most commonly used host is Cape Robin-chat

Cossypha caffra (Skinner 1978). Distribution of the cuckoo in southern Africa is mainly within the combined range of Cape Robin-chat, Heuglin’s Robin-chat C. heuglini and White-throated Robin C. humeralis (Vernon et al. 1997), and it also uses Bearded Scrub-Robin Cercotrichas quadrivirgata (Edwards 1998). In South Africa the Cape Robinchat had 11% (13 of 83) of its nests parasitized in a population in KwaZulu-Natal (Oatley 1970), the region of highest parasitism; overall in southern Africa c. 4.5% of the nests are parasitized (Payne and Payne 1967). In Bioko a fledged young cuckoo was fed by Rufous Ant Thrush Stizorhina fraseri (Pérez de Val et al. 1997) and in southern Cameroon a fledged young was fed by Honeyguide Greenbul Baeopogon indicator (Bates 1930). The eggs are unspotted chocolate brown in southern and East Africa where robin-chat Cossypha species are parasitized, whereas cuckoo eggs that are blue and unspotted or with fine brown spots in East Africa are like those in other hosts especially scrubrobins Cercotrichas spp. In eastern Zaire (Medje) an egg taken from a laying female was pale greenish white, spotted with brown (Chapin 1939, AMNH 159030). In Kenya, at Lake Baringo I found an unspotted blue egg in a nest of White-browed Scrub-robin C. leucophrys (the black-skinned cuckoo nestling hatched on 18 August 1988, Plate 18e,f ), the egg like an unspotted blue egg in the oviduct of a female cuckoo taken 28 May 1967 in Victoria Nyanza (UMMZ 216705). Chocolate eggs in South Africa average 24 ⫻ 18 mm; the Nyanza egg was 22 ⫻ 18 mm. Incubation period is 12–14 days. The nestling cuckoo evicts the eggs and nestlings of its host. At hatching the nestling is naked and black, changing in the next two days to purplish brown, its back is concave and it evicts its nestmates. On the day after hatching the nestling weight is 4.9 g, on day 4 feather sheaths appear on the wing, on day 6 the eyes begin to open, by day 7 the nestling is 20 g, by day 9 it is 30 g (half the weight of a fledgling), by day 10 the eyes are open and the plumage developed, and by the time it is ready to fledge it is 60 g. It fledges in 17–21 days (Skead 1951, Liversidge 1955, Friedmann 1948, 1956, Reed 1969, Newman and Steyn 1970, Newman 1972, Skinner 1978, Rowan 1983, Maclean 1993).

488 Asian Lesser Cuckoo Cuculus poliocephalus

Asian Lesser Cuckoo Cuculus poliocephalus (Latham, 1790) Cuculus poliocephalus Latham 1790, Index Ornithologicus, 1, p. 214. (India) Monotypic. Other common names: Lesser Cuckoo, Small Cuckoo.

Description ADULT: Male, above gray, rump slate gray, wing dark gray, dark rump and tail contrast with the paler back, tail black, rectrices with small white spots along the shaft, white tips, and white notches along the outer edge; the outer feathers T5 are sometimes barred white; underparts, throat and breast light gray, belly white with black bars, under tail coverts white to buff, usually unbarred, bend of wing unbarred white with inner vane of coverts blackish and outer vane white (white lacking in rufous-phase females), under wing coverts white with narrow gray bars. Females have two color phases, one (gray phase) like male but lower breast washed rufous and with narrow blackish bars (lacking in male), the other (rufous phase) above head rufous, back and wing barred rufous and dark brown, rump unbarred rufous to barred rufous and blackish, sides of face and neck rufous, the throat to belly barred black and white, under tail coverts whitish barred with black. Other females are individualistic or intermediate between the gray and rufous phases: one is all rufous above with indistinctly barred back, unbarred rufous rump, and barred rich rufous throat and upper breast (MCZ 12564), another is a laying female with unbarred rich rufous crown and upper back, unbarred gray-brown back, rump and central tail feathers, and white underparts barred dark gray and lacking any rufous (Vogel et al. 2003, AMNH 833647). Both sexes, eye-ring yellow, iris dark brown, bill black, feet tan yellow. JUVENILE: Above slate gray, the blackish crown contrasts with the slate gray back where the feathers often have white edges, white on nape, wing slate sometimes indistinctly barred with buff, face black with white bars; underparts, throat blackish, breast and belly barred black and white, breast (black bars wider than the white bars) separated from the belly (white bars wider than the black bars) by a band of

black, under tail coverts whitish barred black, tail barred black and white with a white tip, bend of wing white with black bars, under wing coverts barred black and white; other juveniles rufous phase, barred rufous buff above, white bars on head and white tips to feathers on back, underparts barred brown and buff or blackish and white. NESTLING: Naked at hatching, skin pink to pinkgray, turning dark gray by day three, corner of mouth yellow, gape lining orange, bill light rose with yellowish upper side, later becoming black with orange at base, feet pink-orange (Netschajew 1977,Yoshino 1999,Tojo et al. 2002. SOURCES: AMNH, BMNH, FMNH, MCZ, MVZ, ROM, SMF, UMMZ, USNM, WAM, ZFMK, ZMUC, ZSM.

Measurements and weights Assam, Sikkim, Nepal and northern India west to Kumaon:Wing, M (n ⫽ 31) 139–159 (149.8 ⫾ 4.4), F (n ⫽ 11) 137–157 (145.8 ⫾ 5.1); tail, M 108–137 (121.73 ⫾ 7.5), F 116–131 (119.4 ⫾ 4.7); bill, M 15.7–19.2 (17.6 ⫾ 0.8), F 16.9–18.7 (17.4 ⫾ 0.7); tarsus, M 15.9–20.4 (18.9 ⫾ 1.4), F 16.4–18.4 (17.9 ⫾ 0.7) (FMNH, ZSM, UMMZ); Japan:Wing, M (n ⫽ 4) 154–161 (155.8 ⫾ 3.6), F (n ⫽ 7) 155–161 (157.1 ⫾ 2.6) (FMNH, UMMZ). Weight, India and Nepal: M (n ⫽ 2) 48–54 (51.0) (ZSM), F (n ⫽ 2) 48–59.5 (53.8) (AMNH, USNM), U 47.9 (Price 1979); Japan: U 42.4–59.6 (54.4) (Enomoto 1941); Zaire: M (n ⫽ 1) 56 (Verheyen 1953). Wing formula, P8⬎ 9 ⱖ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⱖ 10 ⬎ 3 ⬎ 2 ⬎ 1. A subspecies C. p. assamicus Koelz, 1952, described as “smaller and darker” in Assam, Naga Hills, Khasi Hills and Garo Hills, does not differ from cuckoos elsewhere in India.

Field characters Overall length 25 cm. Small cuckoo, above gray, tail with white spots, and underparts barred. In Africa, distinguished from Madagascar Little Cuckoo C. rochii, which also occurs in East and SC Africa, in

Asian Lesser Cuckoo Cuculus poliocephalus 489 part by plumage ((some females have rufous phase in this species, a rufous phase is unknown in C. rochii), and April birds have completed molt, are in fresh plumage and ready to depart to Asia (while April C. rochii are in worn plumage)), and in the hand by size. In Asia, feathered nestlings and juveniles differ from other cuckoos in small size, body weight of feathered 9⫹ day nestlings is 25–30 g (more than 55 g in other cuckoos in Japan); plumage above slaty black, the throat and breast black with fine bars formed by whitish tips of feathers, belly barred black and white, under tail coverts white with black bars (in Oriental Cuckoo C. optatus, juvenile plumage above is dark blackish brown, throat and breast dark blackish brown and the under tail coverts buff ), and by a broad black band across the lower breast (Higuchi and Payne 1986, Higuchi 1998). Juvenile C. poliocephalus are much blacker on the throat and breast than barred juvenile Himalayan Cuckoo C. saturatus. Feathered young also can be told apart by wing length and body size.

Voice Loud husky chattering song “eat your chóky pepper” (La Touche 1931–34, Ali 1962, Ali and Ripley 1969, King and Dickinson 1975,Wells and Becking 1975, Connop 1995, Scharringa 1999), 5 or 6 notes, 1.5–2.8 kHz. In song bouts, variants on this theme are given in a regular sequence, each song 0.8 to 1.2 sec, with 9 songs in 20 sec or more. The first note rises, the second rises then falls, the third note “chóky” has two upper peaks, the fourth is highest with a single peak, fifth is not as high in pitch. Sequences may shift down in pitch over several songs, then rise again, and change in structure, with double peaks in first two notes and single peaks in last three notes (Hwang 1997). Birds also give “quik-quik-quik-quik” calls (Stevens 1925), like the female call in other Cuculus species. They sing at night and long before dawn, more than in the daytime (Martens and Eck 1995).

Range and status N Afghanistan from Gilgit and the frontier to northern Pakistan and India from Kashmir through the Himalayan foothills to Arunachal,Assam in the Khasi Hills and Naga Hills, Bangladesh, Bhutan, China (widespread except in the dry west, occur in southern Tibet,Yunnan and Guangxi),Taiwan, SE Siberia,

Korea and Japan (Hokkaido and Hondo to Kiusiu, Izu Islands and Okinawa) (David and Oustalet 1877, Yen 1933, Kinnear 1934, Ludlow and Kinnear 1937, 1944, Austin 1948, Ali and Ripley 1949, Austin and Kuroda 1953, Gore and Won 1971,Wells 1972,Wells and Becking 1975,Wild Bird Society of Japan 1980, Flint et al. 1984, Cheng 1991, Roberts 1991,Vuilleumier 1993, McWhirter et al. 1996, Grimmett et al. 1999, Yoon 2000). In southeast Asia they breed in NW Thailand, northern Laos and Tonkin; further south in Vietnam they sing: their status is uncertain (Stanford and Ticehurst 1939, Robson 2000a, AMNH). Migratory, cuckoos occur in their breeding range in Siberia, China and Japan from May to September, and are passage migrants in Burma (Maymyo, Osmaston 1916; Karen Hills, Smith et al. 1943), northern Thailand (Chiang Mai, Wells and Becking 1985; USNM), southern China (La Touche 1930–34, MacKinnon and Phillipps 2000), Okinawa and Nansei-shoto (Kuroda 1925, McWhirter et al. 1996). Resident in Taiwan, numbers increase in winter when northern birds are there (Hachisuka and Udagawa 1951); and resident in Hainan (Guangdong et al. 1983, Cheng 1991). They migrate at night and are attracted to lights (Stevens 1925). In southern India they occur in the Eastern and Western Ghats in passage in October and May (Price 1979, BMNH).They are unknown in winter in southeast Asia south of 19° N except rarely in the Andaman Is and Sri Lanka (Kotagama and Fernando 1994). Cuckoos appear in autumn and spring migration in the Indian Ocean on Île Amsterdam, the Seychelles, and the Mascarenes on Réunion and Mauritius (Staub 1976, Becking 1981, Roux and Martinez 1987). Cuckoos winter in East Africa from November to April, are seen mainly near

490 Asian Lesser Cuckoo Cuculus poliocephalus the coast, in Tanzania from the Uluguru Mts to Mikindani (ZMUC, ZSM).There are two records in N Somalia in November (Ash and Miskell 1998), and a few records in southern coastal Kenya and in coastal and southern Tanzania (Zimmerman et al. 1996, Stevenson and Fanshawe 2002), southern Zaire (once, Upemba NP,Verheyen 1953), Zambia east of 31° E (Benson et al. 1970,Aspinwall and Beel 1998), Malawi (Benson and Benson 1977), Zimbabwe, Mozambique and South Africa (Irwin 1981, 1982, 1988). A bird ringed in coastal Kenya in November was recovered inland in December. In Africa, flushed birds may give a loud staccato rattle (Stevenson and Fanshawe 2002); April birds are migrants, usually silent and fat (Lewis and Pomeroy 1989).

Habitat and general habits Montane in much of its breeding range. Forests, broad-leafed and pine, scrub, second growth, in Ussuria in SE Siberia and NE China in forests of birch, hornbeam, maple and oak, in Himalayas in Nepal mainly in closed, dense evergreen broad-leaved cloud forests to oak Quercus, Rhododendron and coniferous forests of the subalpine zone. In India they occur in the breeding season mainly from 1400 to 3050 m (Baker 1927,Ali 1942), in Nepal and Bhutan mainly between 1500 and 3500 m (Inskipp and Inskipp 1985, Martens and Eck 1995, Grimmett et al. 1999), in northern Burma between 2300 and 3600 m (Cranbrook, in Kinnear 1934, Stanford and Ticehurst 1935, Smythies 1940); in beech forests at 1450– 1500 m in Guangxi, China (Vuilleumier 1993), and in lowlands to 2300 m in Japan (Brazil 1991). In Africa, they live in forests and woodland (Irwin 1988); this is the only cuckoo in pine plantations (Benson and Benson 1977), similar to its breeding habitat in northern India and eastern Asia. In Africa usually silent, but sometimes call in E Africa before migration in spring.

Food Insects, mainly caterpillars (geometrids, noctuids), and beetles, mantids and small bees (Baker 1927, Dement’ev and Gladkov 1966).

Breeding In the plains of India the song starts in April, in the Himalayas eggs are laid late May to July, with the

late record a fledgling just out of the nest in August (Baker 1927, 1934, Ludlow and Kinnear 1937, Ali and Ripley 1969, Becking 1981, Martens and Eck 1995). In Vietnam, a rufous-phase female had an egg nearly ready to lay (9 mm) in May (AMNH 833647). Brood-parasitic. In southeast Siberia in Ussuria, the hosts are Bush Warbler Cettia diphone and other small songbirds (Neufeldt 1968, 1971, Netschajew 1977, Knystautas and Sibnev 1987). In Japan, the main host on Honshu is Bush Warbler Cettia diphone and cuckoos are also reared by Winter Wren Troglodytes troglodytes, while on Izu Islands where other cuckoos are absent the Little Cuckoo successfully uses both these two hosts and Ijima Willow Warbler Phylloscopus ijimae, Western Crownedwarbler P. occipitalis, Middendorff ’s Grasshopper Warbler Locustella ochotensis, and three other species (Higuchi 1998). The nest of Bush Warbler has its entrance at the side, the female cuckoo pushes her head into the hole, finds the host egg, then pushes her body into the nest to lay her own egg (Yoshino 1999). In Kashmir a recent host record for a nestling cuckoo is Tickell’s Leaf-warbler P. affinis and another is Blyth’s Leaf-warbler P. reguloides (Price and Jamdar 1991, Marchetti 1992). In India other hosts are Brownish-flanked Bush-warbler Cettia fortipes and Pale-footed Bush-warbler C. pallidiceps, and perhaps other small warblers and babblers including wrenbabblers Pnoepyga and shortwings Brachypteryx; only in bush-warbler records did the cuckoo egg hatching enable identification (Baker 1934, 1942, Ali 1962, Becking 1981). Old breeding reports in Malay Peninsula and Java (Hoogerwerf 1949, Madoc 1956a,b) were of another cuckoo, Cuculus lepidus. Eggs, in Siberia the unmarked brown eggs are 17.5 ⫻ 14.2 and 19.0 ⫻ 15.3 mm (Neufeldt 1971), in Ussuria the eggs are unmarked brown, 19.1 ⫻ 14.5 mm and 2.8 g when fresh, in Japan the eggs are unmarked brown, 21.7 ⫻ 15.7 mm (Higuchi 1989, 1998). In India unmarked brown eggs were taken from the oviduct of laying females and looked like cuckoo eggs in nests of bushwarblers (Baker 1906). Other eggs that Baker (1906, 1934) attributed to this cuckoo were (1) red, and (2) unmarked white, both matching the eggs of suspected host species. An egg from a laying female

Sulawesi Cuckoo Cuculus crassirostris 491 cuckoo was white, like cuckoo eggs from nests of crowned-warbler (Baker 1906); this warbler usually rejects eggs that are unlike its own in color or size (Marchetti 1992).The cuckoo female removes 1–2 host eggs from the nest. The incubation period is unknown; the young cuckoo hatches two to zero days before the host hatches. The young cuckoo evicts the host eggs, beginning to evict a few hours after it hatches and retaining the behavior for 2–3 days. The nestling takes 7–10 min to evict the egg and takes a 2 min

rest until it evicts the next egg, and it can evict eggs from a nest that is much deeper (6.5 cm) than its own body. After it evicts the host eggs, it crouches in the bottom of the nest, the head held backwards on the back, the open bill held vertical.The nestling grows to 5 g by the second day, 12 g by the fourth day and 20 g by the sixth day, and it can perch on the nest rim by 12 days. Nestling period is 18–20 days, and the young bird fledges at 40 g. In Ussuria, 4 of 25 nests of Bush Warbler were parasitized by this cuckoo (Netschajew 1977).

Sulawesi Cuckoo Cuculus crassirostris (Walden, 1872) Hierococcyx crassirostris Walden, 1872, The Annals and Magazine of Natural History, (4), 9, 305. (northern Celebes) Monotypic. Other common names: Sulawesi Hawk-cuckoo.

Description The variation in plumage is considerable and the sequence of plumages has not previously been described. No short-tailed juveniles are known in museum collections, but the white-crowned birds are juveniles, insofar as in other cuckoos particularly in the closely related Indian Cuckoo C. micropterus the white-headed plumage is of the juvenile plumage, and in C. crassirostris that are otherwise in juvenile plumage the black feathers of the crown and buff feathers of the throat appear, indicating a transition from juvenile plumage to subadult plumage. ADULT: Sexes alike (?), crown gray, back graybrown to brown, wing coverts and flight feathers dark brown, tail banded black and rufous, T1 and T2 dull rufous with a black subterminal band and a small white tip, T3 and T4 dull rufous with small white spots along the shaft and indistinct black bands, T5 with black and white bars and white notches on outer vane, head and cheeks gray; underparts white with broad black bars, a broad black band across the upper breast, under tail coverts white, bend of wing white, under wing coverts white, wing with base of primaries P1–P9

white showing a broad white band, P10 not banded; iris dark brown, bill brown to black above, below green base, black at tip, feet yellow. Female plumage is known from a specimen (USNM 112699) without sex written on the field label; later on “F?” was added to the label; this bird is similar in plumage to male cuckoos. SUBADULT: Head black with some white, back bright rufous, wing rufous, tail as in adult, underparts buffy white, throat washed buff, breast and belly with a few black spots and bars, especially on the upper breast where the black pectoral blotches form a nearly continuous band across the breast. JUVENILE: Head all white with a few black feathers appearing on the nape, back bright rufous, wing rufous, tail as in adult with more white spots along the shaft of T1, underparts uniformly white from chin to under tail coverts, with buffy feathers appearing on the throat. NESTLING: Undescribed. SOURCES: AMNH, BMNH, MSNG, MZB, RMNH, SMTD, UMMZ, USNM, ZMA, ZMB.

Measurements Wing, M (n ⫽ 10) 200–212 (206.1 ⫾ 2.8), F (n ⫽ 1) 198, U (n ⫽ 8) 200–211 (203.8 ⫾ 4.0); tail, M 162–173 (168.2 ⫾ 2.4), F 160, U 146–180 (169 ⫾ 20.9); bill, M 25.0–28.2 (27.3 ⫾ 1.6), F

492 Indian Cuckoo Cuculus micropterus 26.0, U 24.7–27 (25.5 ⫾ 0.9); tarsus, M 22.8–26.7 (24.5 ⫾ 1.8), F 27.7, U 25.5–30 (26.7 ⫾ 2.0) (AMNH, BMNH, MSNG, RMNH, SMTD, UMMZ, USNM, ZMA, ZMB). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6(or 9 ⫽ 6)⬎ 5 ⬎ 4 ⬎ 3 ⬎ 10 ⬎ 2 ⬎ 1.

Field characters Overall length 34 cm. Large cuckoo with a gray head, dark brown back, black and rufous banded tail, and heavily barred underparts from chin to belly.

Voice Two far-carrying mellow notes at a low pitch, the first note higher (1 kHz) than the second (0.9 kHz), “ho-oo” (“cuck-oo”), or three notes “ho-oo-oo” or “kokokuk” with each note lower than the one before, or four notes “kO kO ku (ku),” the last note lower and softer; song duration c. 1.0 sec. Also gives a “dong, dong” (Heinrich, in Stresemann 1940, Watling 1983, Van den Berg and Bosman 1986, Rozendaal and Dekker 1989, Gibbs 1990, Holmes and Phillipps 1996, Coates and Bishop 1997; Ben King recording). Calls persistently at night and near dawn and occasionally on overcast mornings.

ent subadult plumage suggest a “Hierococcyx” hawkcuckoo, but the cooing call is more like that of Cuculus, e.g. Common Cuckoo Cuculus canorus.

Habitat and general habits Primary montane forest, tall second growth hill forest and forest edge, they live in the canopy, perch upright, often in leafy midstory, near the trunk of a large tree; they occur from 500 to 1400 m.

Food Insects, including termites (Stresemann 1940).

Range and status Mountains of northern and central Sulawesi. Resident. Uncommon to rare, along a transect there are fewer than 1/km where they occur.They are heard in northern Sulawesi at S. Moinakom, Edwards’ camp, Clark’s camp, Hogs Back and Wein camp in Cumoga-Bone NP, northern Sulawesi.They are shy, inconspicuous and difficult to observe (Rozendaal and Dekker 1989). The broad wing and the appar-

Breeding Their songs are heard at Lore Lindu from May to August (White and Bruce 1986). Presumably a brood-parasite.The eggs and other details of breeding are unknown. A young bird perhaps of this species begged with “coos” (“girren”) from a Haircrested Drongo Dicrurus [hottentottus] leucops (Heinrich, in Stresemann 1940).

Indian Cuckoo Cuculus micropterus Gould, 1837 Cuculus micropterus Gould, 1837, Proceedings of the Zoological Society of London, 1837, p. 137 (Himalayas) Polytypic. Two subspecies. Cuculus micropterus micropterus Gould, 1837; Cuculus micropterus concretus (S. Müller, 1845).

Description ADULT: Sexes similar, male, above dark slate gray, back brownish gray, wing slate gray, tail gray with a broad black subterminal band, the inner feather with alternating small black and white spot on shaft

Indian Cuckoo Cuculus micropterus 493 T1, alternating black and white spots along feather shafts T2 to T4, white bars on outer tail feathers T5; underparts, throat and upper breast pale gray bordered sooty on posterior side of breast band, belly white with widely spaced black bars (black bars 2–3 (to 4) mm, white bars 6–7 mm), under tail coverts white with black bars, bend of wing white, under wing coverts whitish with a few broad black bars, a gray band formed by the greater wing coverts, and a broad whitish band across the secondaries and primaries (all but the outer three); eye-ring gray to yellow, iris light brown to reddish brown, bill black, greenish below, feet yellow; female, similar but throat pale gray and breast brownish, tail with brown in the gray bands, belly with black bars narrower than in male. JUVENILE: Above, crown and nape nearly all white (feathers black at base, the terminal 3–4 mm white), the back, rump and upper tail coverts slate gray broadly tipped and barred with rufous and white, wing flight feathers dark brown tipped with rufous buff, tail as in adult but more rufous, face with broad black band below eye, underparts with chin and throat white giving a masked appearance; underparts whitish buff spotted and barred with black. The extent of conspicuous pale plumage on the juvenile is highly variable with white or buffy patches, and the head is sometimes almost entirely pale. NESTLING: Naked at hatching, skin yellowish pink; gape flange yellow, mouth lining orange-red. SOURCES: AMNH, ANSP, BMNH, BPBM, CM, DMNH, FMNH, MCZ, MZB, RMNH, ROM, UMMZ, USNM,YPM, ZRC.

Subspecies and history Cuculus micropterus micropterus Gould, 1837; larger and paler; northern China and SE Russia, Mongolia, Korea, N Pakistan, Kashmir and Himalaya foothills, south through India, east to Nepal, Bangladesh, Sri Lanka, Burma, Bay of Bengal islands, Thailand, Laos, eastern China, Hainan and the Philippines; winters south to Malay Peninsula,Andaman Islands, Greater Sundas (Sumatra, Java, Borneo, Bangka) and the Philippines;

Cuculus micropterus concretus S. Müller, 1845; smaller and darker; Malay Peninsula (from Trang in extreme southern Thailand), Sumatra, Java and Borneo. Birds in the Amur area of Russia and China north of Shandong average larger than birds in southern Asia (Vaurie 1965). Swinhoe (1870) collected one in northern China and described it as Cuculus Michieanus, wing 8.3 inches (211 mm). Although northern birds are larger on average, some birds in northern China are small and some birds in India during the breeding season are large. Birds in the Philippines are similar in plumage, size and song to birds in mainland Asia. Koelz (1954) described C. m. fatidicus from Cachar, India, as darker than birds in the Himalayas, but birds in this region are not consistently darker than birds in Assam (UMMZ), and dark birds occur elsewhere as on Hainan. The large wings (220 mm) reported for live migrants in the Malay Peninsula (Wells 1999) are much larger than seen in any museum specimens.

Measurements and weights C. m. micropterus, Assam, N India and Nepal: Wing, M (n ⫽ 18) 188–205 (195.3 ⫾ 5.7) (n ⫽ 5 over 200, 4 April to 15 May), F (n ⫽ 7) 189–203 (193.6 ⫾ 6.8); tail, M 140–167 (158.6 ⫾ 7.2), F 152–161 (157.0 ⫾ 5.9); bill, M 23.5–26.4 (25.8 ⫾ 0.9), F 24.2–27.0 (25.3 ⫾ 1.2); tarsus, M 19.1–24.6 (21.2 ⫾ 1.9), F 19.1–22.0 (22.0) (AMNH (FF), FMNH (FF), UMMZ); SE Russia (Ussuria, Amur), Korea, and China north of 30°N: Wing, M (n ⫽ 20) 189–215 (204.6 ⫾ 6.3), F (n ⫽ 7) 198–214 (203.8 ⫾ 6.3) (ANSP, FMNH, MCZ); Philippines: Wing, M (n ⫽ 10) 187–206 (194.9 ⫾ 8.5), F (n ⫽ 5) 185–200 (193.2 ⫾ 6.1) (AMNH, BPBM, DMNH, FMNH, USNM); C. m. concretus: Wing, M (n ⫽ 10) 151–168 (162.9 ⫾ 5.9), F (n ⫽ 6) 151–170 (160.8 ⫾ 9.0) (AMNH, BMNH, RMNH,YPM). Weight, C. m. micropterus, Russia: M (n ⫽ 6) 112–129 (119.0), F (n ⫽ 1) 119 (Neufeldt 1966); India: M (n ⫽ 3) 100.6–109 (103.9), F (n ⫽ 1) 108 (MCZ); Malay Peninsula (autumn): (n ⫽ 17)

494 Indian Cuckoo Cuculus micropterus 77.5–121.8; Philippines: M (n ⫽ 5) 98–121.9 (105.4) (BPBM, FMNH); Sumatra: M (n ⫽ 1) 106 (ANSP); C. m. concretus: U (n ⫽ 2) 63.0–69.3 (66.2) (Wells 1999). Wing formula, P8 ⬎ 9 ⬎ 7 ⬎ 6 ⬎ 5 ⬎ 10 ⬎ 4 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 32 cm. Cuckoo with gray upperparts, underparts with broad black bars, and a barred tail with a broad black subterminal band and white tip, and more conspicuous white bars and spots on the other tail feathers. Differs from other cuckoos in broad black subterminal band and white tip on the tail, and broader black bars on the lower breast and belly. Juvenile has broad white markings on crown, and a black face mask with white chin and throat, distinct from that of other cuckoos in its range.

Voice Loud flute-like whistle, four notes, three nearly on one pitch, or alternating high and low, the first note rising in pitch, the first or the third sometimes the highest, the fourth soft and low, the call an inflected “orange-pekoe”, “crossword puzzle” or “one more bottle”. In China the song sometimes rises at the end (“la-sol-sol-mi”, David and Oustalet 1877). In the Indian subcontinent in Kangra Valley, the cuckoo is known as the soul of a dead shepherd who is eternally searching, “where is my sheep?” (Whistler 1926).The song recalls the introduction of Beethoven’s Symphony no. 5, off-key.The first three notes are at 1.4 kHz, the last note at 1.3 kHz, and the full song is 0.7–0.8 sec long. Songs in recordings are nearly identical in India, Japan, Korea, SE Russia and Nepal, and songs in verbal descriptions are similar in China, Burma, Malay Peninsula and the Greater Sundas (Swinhoe 1870, Osmaston 1916, Hoffmann 1940, Smythies 1940, 1960, Neufeldt 1966, King and Dickinson 1975,Wells and Becking 1975, Medway and Wells 1976, Wild Bird Society of Japan 1982, Flint et al. 1984,White 1984, MacKinnon and Phillipps 1993, Martens and Eck 1995, Ali 1996, Davison 1997, Hwang 1997, Grimmett et al. 1999, Scharringa 1999,Wells 1999) and in the Philippines (R. S. Kennedy, tape recording).The male calls per-

sistently in the breeding season. Another call is a long crescendo, given in excitement. Female gives a rushed bubbling call or chuckle (Neufeldt 1966).

Range and status Widespread through the Indian subcontinent except in arid western areas of Rajasthan and North Gujarat, they occur mainly in the Himalayas from Punjab and Himachal Pradesh east to Nepal, Bhutan and Arunachal Pradesh, Assam including Meghalaya south to Lushai Hills, eastern Bangladesh (Chittagong), and south through the peninsula including Maharashtra and the western Ghats, also in Sri Lanka, southeast Asia (at least in the northern parts), and the Bay of Bengal islands (Davidson 1886, Ali 1946, 1953, 1996, Grubh and Ali 1947, Ali and Ripley 1969, Roberts 1991, Grimmett et al. 1999). In China they are widespread except in the high and dry northwestern half of the country (Cheng 1991), in Burma they are in the Chin and Shan hills (Smythies 1986), in Thailand in the northwestern hills, and they are in Laos (FMNH), Cambodia (Thomas and Foote 2003) and Vietnam (Annam, BMNH). In Russia they are local in the southeast in the Amur River region and Ussuria (Flint et al. 1984). Summer resident in temperate Asia, passage migrant in Korea where birds remain and sing in June (Gore and Won 1971, Tomek 1985, Fiebig 1995, Hwang 1997,Yoon 2000), accidental in Japan (Brazil 1991). Northern birds winter in tropical southern Asia and on islands from the Andamans and Sumatra to Java and Borneo, one record in the

Indian Cuckoo Cuculus micropterus 495 Moluccas (Ternate). Southern populations are resident (Guangdong et al. 1983, Mees 1986,White and Bruce 1986, van Marle and Voous 1988, Cheng 1991, Coates and Bishop 1997). In the Philippines, most records are from late October to May and may be of winter visitors. Nevertheless, cuckoos sing in the Philippines in winter (Kennedy et al. 2000) and breeding though previously overlooked is indicated by a growing juvenile on Luzon. On the Malay Peninsula, cuckoos occur as residents, nonbreeding winter visitors and passage migrants. Large birds, C. m. micropterus, occur in all months but are less common in December and January, suggesting that they move southward, while early winter records on small islands in Malacca Straits and a lighthouse strike indicate southward passage (Wells 1999). Uncommon in Sri Lanka (Kotagama and Fernando 1994). In Sumatra, breeding C. m. concretus occur in mountains in Utara, Barat and Bengkulu and on the Batu Islands; wintering C. m. micropterus occur in Aceh and Utara and in the Malacca Straits, Lingga Archipelago, Bangka, Belitung and Nias (van Marle and Voous 1986, Mees 1986). In West Java, both forms are local and uncommon (Kuroda 1936). In Borneo, resident C. m. concretus are more numerous than wintering migrant C. m. micropterus (Smythies 1957, 1981, Sheldon et al. 2001). Breeding density, in northern India as high as 1.5 calling birds per mile2 (c. 0.5 birds per km2) (Fleming 1967).

Habitat and general habits Forests both deciduous and evergreen, second growth, dry country, thick scrub (Neufeldt 1966). They occur in montane forests and wooded valleys, and in wooded parks and groves near villages more than do other cuckoos in Nepal. Lowlands to hills at 2000 m, sometimes to 2800 m, in Nepal to 3700 m, mainly 500–1900 m. In the Himalayas they breed at lower altitudes than other cuckoos (Stevens 1925, Fleming 1967, Inskipp and Inskipp 1985, Martens and Eck 1995).Tropical C. m. concretus occur in tall forest canopy, at plains level and slopes to 800 m in Malay Peninsula (Wells 1999) and in Melaleuca and peat swamp forest along rivers in Kalimantan (van Balen and Prentice 1997). C. micropterus feed in high forest canopy, take flying termites in the air, and also feed on the ground.

Food Insects, mainly hairy caterpillars (Siberian moth Dendrolimus sibiricus, a pest on larches; tussock moth Dasychira albodentata) and butterflies, grasshoppers, mole crickets, click beetles, carpenter ants, flying termites; fruit (Ali 1962, 1996, Dement’ev and Gladkov 1966, Neufeldt 1966, Smythies 1981, Sody 1989, Roberts 1991,Wells 1999).

Displays and breeding behavior Males in breeding season are spaced in forest, countersing, and are territorial (Wells 1999). In China, Herbert Lynes noted on a label,“There were about four males calling (the 4-syllabled note) and one or more females bubbling like female Cuculus canorus. After calling I saw one of the males squat-droop wings, up tail, a little above her—and sway body from side to side about legs as a pivot” (BMNH 1934.1.1.3537). In Amurland, males sing alternately except when a female is nearby, then the males call together.A female cuckoo used an area of 50–100 ha where there were 25–30 pairs of host shrikes. During the laying season, cuckoos live in pairs. The female removes and eats an egg from the host nest (Neufeldt 1966).

Breeding In Amurland they lay in June (Neufeldt 1966), in northern China from May to July (Shaw 1938, Hoffmann 1950, BMNH), in India they call March to July or August and there are egg records in May and June (Inglis et al. 1920, Baker 1934, Price 1979, Ali 1996; Punjab, 2 May 1933, UMMZ 82672), in Sri Lanka they call January to May (Kotagama and Fernando 1994), in Burma from January to June (Smythies 1940), in Malay Peninsula from January to August, mainly March to June (Wells 1999, MZB), in Java eggs are laid in April, June and December (Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967) and adults call the rains during from January to May (Mees 1949), and in Sabah the song is heard mainly from October to April (Sheldon et al. 2001). In the Malay Peninsula a female with an oviduct egg was taken in April, and fledglings and grown juveniles date from June to early September (Wells 1999). In the Philippines the cuckoos sing in winter in December and January

496 Madagascar Lesser Cuckoo Cuculus rochii (Kennedy et al. 2000, R. S. Kennedy, pers. comm.), and in Luzon a juvenile with a white crown and short, growing wing and tail (FMNH 265472) was taken on 17 May 1961. This young bird provides the first evidence for this cuckoo breeding in the Philippines. Brood-parasitic. Hosts, in Russia they lay in nests of Brown Shrike Lanius cristatus (Neufeldt 1966), in China they use Black Drongo and Azure-winged Magpie Pica cyanea and probably Brown Shrike (La Touche 1931–34, Shaw 1938, Hoffmann 1950). In India their eggs were in nests of drongos, fledged young cuckoos were fed by Black Drongo Dicrurus macrocercus and Ashy Drongo D. leucophaeus (O’Donel 1936, Jones 1941, Storrs 1944, Becking 1981), and a fledged young cuckoo was fed by Streaked Spiderhunter Arachnothera magna (Gammie 1877). In the Malay Peninsula at Selangor fledglings were fed by Black-and-yellow Broadbills Eurylaimus ochromalus (Wells 1999), and in Java an egg was found in a nest of Ashy Drongo Dicrurus leucophaeus and a young cuckoo was fed by Greater Racket-tailed Drongo Dicrurus paradiseus (Hellebrekers and Hoogerwerf 1967, Becking 1981; RMNH). Eggs are whitish with reddish-brown markings in Lower Amur in Russia, like eggs of the host shrikes (Neufeldt 1966), and are similar in China (oviduct egg, La Touche 1931–34). Eggs are either white with reddish-brown markings or unmarked blue in India (Baker 1906, 1907, 1908a, 1934, 1942). The

blue egg identifications of Baker, based on eggs in nests in Murree, Pakistan, by Rattray (Baker 1906), an unmarked oviduct egg (Baker 1934: 347) and a fresh egg in a nest of Asian Paradise-Flycatcher Terpsiphone paradisi found when a cuckoo flew from the nest (Baker 1934: 347) were questioned by Becking (1981). Eggs are whitish with reddishbrown markings in the Malay Peninsula (Colebrook-Robjent 1978) like eggs of the broadbill, and eggs are similar in Java (Hellebrekers and Hoogerwerf 1967). Eggs are 25 ⫻ 19 mm in Lower Amur (Neufeldt 1966), 23.6 ⫻ 17.7 mm in Java (Hoogerwerf 1949). Incubation period is 12 days (14 days in shrikes).The nestling cuckoo is 4.7–4.9 g at hatching, it evicts host eggs or chicks on day 3 after hatching. Through day 2 the nestling is naked, on day 3 minute hair-like down appears in the site of the flight feathers; on day 4–5 the feathers begin to appear, the tips of undeveloped feathers beige or pale yellow, the eyes open on day 7; and the chick is well feathered by day 13–14. Nestling perches on edge of nest on day 18, the normal nestling period is 21 days, and a young cuckoo flew well at 30–40 days (Neufeldt 1966, Makatsch 1971). In India a fledgling reared by Black Drongo had a call like a young drongo (Osmaston, RMNH 27040). Frequency of brood parasitism, 3 of 50 nests of Brown Shrikes in Amur had a cuckoo egg or nestling (Neufeldt 1966). Birds molt in winter, and the young birds often retain some juvenile feathers in their first breeding season (Neufeldt 1966).

Madagascar Lesser Cuckoo Cuculus rochii Hartlaub, 1863 Cuculus rochii Hartlaub, 1863, Proceedings of the Zoological Society of London, 1862, p. 224. (Madagascar). Monotypic. Other common names: Madagascar Cuckoo.

Description ADULT: Sexes alike or similar, above slate gray, wing and tail darker gray to black (little contrast with back), the tail feathers with white spots on outer edge (especially outer feathers T4,T5), small white spots along shaft and white tip, below the

throat and breast light gray, belly white with black bars, under tail coverts whitish to buffy, unbarred or with a few bars, bend of wing with a few whitish feathers (barred in rufous juveniles), under wing coverts white with narrow black bars. Females variable in plumage, some gray like males, others with a rufous wash on breast and rufousgray throat, but a distinct barred rufous phase is unknown; eye-ring yellow, iris brown, bill black with yellowish base, bill broader at base than in C. poliocephalus, feet yellow.

Madagascar Lesser Cuckoo Cuculus rochii 497 JUVENILE: Above dark brown barred light rufous, the feathers tipped white, the rump and upper tail covert feathers black with whitish tips, wing brown with rufous bars and white tips, tail dark gray, face barred blackish and white; underparts, throat blackish (feathers white at base and tip), breast and belly to under tail coverts white with narrow black bars NESTLING: Undescribed at hatching.The feathered nestling has a yellow gape flange and a red inner mouth and tongue (Hawkins and Goodman 2003). SOURCES: AMNH, BMNH, MNHN, SMTD, USNM, ZSM, ZFMK, ZMB.

Measurements and weights Madagascar: Wing, M (n ⫽ 20) 155–175 (166.1 ⫾ 4.4), F (n ⫽ 7) 153–168 (159.7 ⫾ 5.6); tail, M 124– 152 (140.8 ⫾ 8.6), F 137–151 (139.8 ⫾ 2.5); bill, M 16.3–19.4 (18.7 ⫾ 2.2), F 16–19 (17.5 ⫾ 0.9); tarsus, M 15.3–19.0 (16.6 ⫾ 0.8), F 15.5–18 (16.7 ⫾ 0.8) (AMNH, BMNH, USNM). Weight, M (n ⫽ 2) 64–65 (64.5) (Benson et al. 1976). Wing formula, P8 ⬎ 9 ⱖ 7 ⬎ 6 ⬎ 5 ⬎ 4 ⱖ 10 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 26 cm. Small cuckoo, above gray, tail with white spots, breast gray and belly barred black and white. Distinguished from Asian Lesser Cuckoo C. poliocephalus, which also occurs in East and SC Africa in different seasons, by song, longer tail, and in part by plumage (a small rufous adult cuckoo in this region is Asian Lesser Cuckoo; C. rochii juvenile upperparts very dark brown, barred rufous and white, underparts uniformly barred (throat and breast nearly black in C. poliocephalus). In Africa, birds are in worn plumage in April when Asian Lesser Cuckoo are in fresh plumage. In the hand, it is larger than Asian Lesser Cuckoo and the plumage is darker gray.

Voice Four evenly spaced notes, “ho-ho-ho-hu”, or “chew chew chew choo”, 0.7 to 1.2 kHz, each

note falling in pitch, the first three notes higher, the fourth lower; the song is lower in pitch than that of Asian Lesser Cuckoo C. poliocephalus. Calls for hours at a time, often at night, from high perches in forest or forest edge. Female gives a loud series of single notes “pee pee pee pee pee” which on playback attracts singing males (Milon et al. 1973, Dowsett and Dowsett-Lemaire 1980, Becking 1988, Goodman et al. 1997, Randrianary et al. 1997, Morris and Hawkins 1998, Chappuis 2000).

Range and status Madagascar. Migratory. Common and widespread in Madagascar (Rand 1936, Milon et al. 1973, Goodman et al. 1997, Morris and Hawkins 1998). They arrive and sing in September and October, later in the dry southwest where they are less common. After breeding, most birds move from Madagascar northwest to Africa in Tanzania, Kenya, Uganda, E Zaire, Rwanda, Burundi and Zambia (E of 28° E) to Malawi, Zimbabwe and northern Mozambique, and occasionally in Mauritius. In Africa they generally appear from April to September ( Jackson 1938, Benson et al. 1970, Becking 1988, Irwin 1988, Louette and Herroelen 1993, Aspinwall and Beel 1998, Rossouw and Sacchi 1998). There is one record in South Africa in December, a bird singing at Mala Mala, Transvaal (Hockey 1992, Chappuis 2000, where called “C. poliocephalus” ).

498 African Cuckoo Cuculus gularis

Habitat and general habits

Breeding

Nearly any area with any trees, from forest (including eucalyptus plantations and gallery forest) to scrub woodland and spiny forest, and in marshes; not noted in littoral forest. They occur from sea level to 1950 m, most numerous at low elevations and rare above 1500 m, a few in heathlands at 2200 m. In nonbreeding season in Africa, they occur in forests and woodland to acacia bush.They sometimes join mixed-species feeding flocks (Rand 1936, Milon et al. 1973, Langrand 1990, Goodman and Putnam 1996, Zimmerman et al. 1996, Goodman et al. 1997, 2000, Hawkins et al. 1998, Hawkins and Goodman 1999).

They sing October to December, egg records August–April (Rand 1936, Milon 1959, Morris and Hawkins 1998). Brood-parasitic. Hosts, Madagascar Cisticola Cisticola cherina (10 of c. 210 clutches were parasitized) and Common Jery Neomixis tenella (1 of c. 50 clutches), Grey-crowned Tetraka Phyllostrephus (Xanthomyzis cinereiceps), Madagascar Paradise Flycatcher Terpsiphone mutata (1 of 8 clutches), and Madagascar Sunbird Nectarinia souimanga. Eggs are whitish, often tinged with pink and spotted with reddish brown; the color and pattern do not closely match the eggs of the host species; c. 18 ⫻ 14 mm, 1.8 g (Rand 1936, Milon 1959, Benson et al. 1976, Hawkins and Goodman 2003). Incubation and nestling periods are unknown.

Food Insects, hairy caterpillars (Rand 1936).

African Cuckoo Cuculus gularis Stephens, 1815 Cuculus gularis Stephens, 1815, in Shaw’s General Zoology, 9, p. 83. [Camdeboo, Cape Province] Monotypic.

Description ADULT: Sexes similar, above, head and back dark ashy gray, wing ashy gray, tail blackish barred and tipped with white, no subterminal black band, rectrices black in center with white spots along the feather shaft and gray along the edge, the central rectrices T1 and T2 without spots, outer rectrices T3 to T5 with white spots along the edge, often with white bars on outer tail feathers T5, tail appears barred from below; underparts, chin to breast ashy gray, belly and under tail coverts white with narrow black bars, bend of wing white with dark gray bars; eye-ring yellow, iris yellow, bill yellow above from base to 3–4 mm in front of nostril and most of the lower mandible with black tip; female is sometimes slightly barred or rufous on throat or breast, no rufous phase. JUVENILE: Above ashy gray with irregular darker gray bars (not brown), tail with broad white spots,

the spots joined into white bars on the outer feathers T5; underparts whitish with blackish bars on throat, the breast to under tail coverts white with thin black bars, tail appears barred from below; iris olive brown, feet orange yellow. NESTLING: Naked at hatching, the skin dark purple or blackish, the gape bright orange, the feet yellow. SOURCES: AMNH, ANSP, BMNH, BWYO, CM, FMNH, MCZ, MVZ, NMK, ROM, UMMZ, USNM,YPM, ZSM.

Measurements and weights Wing, M (n ⫽ 9) 204–226 (215.3 ⫾ 8.4), F (n ⫽ 9) 197–209 (205.3 ⫾ 5.1); tail, M 143–166 (152.9 ⫾ 10.8), F 141–156 (147.0 ⫾ 5.5); bill, M 20.7–23.6 (21.6 ⫾ 0.5), F 19.0–22.4 (20.7 ⫾ 1.2); tarsus, M 20.8–22.3 (21.5 ⫾ 0.6), F 20.5–22.8 (21.7 ⫾ 1.0) (AMNH, UMMZ, ZSM). Weight, M (n ⫽ 1) 97, F (n ⫽ 1) 110 (UMMZ, ZSM). Wing formula, P8 ⬎ 7 ⱖ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 10 ⬎ 3 ⬎ 2 ⬎ 1.

African Cuckoo Cuculus gularis 499

Field characters Overall length 32 cm. Large gray cuckoo, the lower breast and belly with black bars, under tail coverts white either unbarred or with black bars, tail with distinct white bars, and bill mostly yellow (dark at tip). In breeding season, distinguished by its slightly rising song (song descends in Common Cuckoo C. canorus). In the hand, the bill at base is more broad and deep than in C. canorus. In Africa some C. canorus have bill yellow at the base but not forward of the nostril; these birds have a narrow bill. Juveniles differ from juvenile C. canorus in the lack of brown in the plumage and in the broader white spots and bars in the tail ( Jackson 1938, Payne 1977a).

Voice Male song a loud burry whistle “coo-coó”, second note higher and louder; female a bubbling “kwikkwik-kwik!”. Song is similar throughout Africa (North 1958, Chappuis 1974, 2000, Payne 1986, Irwin 1988, Stjernstedt 1993).The distinct song was first noted in 1920 in Darfur, Sudan (Lynes 1925) and 1923 in northern Nigeria, and differs from that of the northern Common Cuckoo which has a louder first note (Bates 1924, Bannerman 1933).

Range and status Africa from Senegal and The Gambia, southern Niger, northern Nigeria, Sudan, Ethiopia and Kenya, south to Namibia, Botswana, Zimbabwe, Mozambique, Transvaal and Natal, occasionally south to Grahamstown-East London area of E Cape; absent from the dry southwest and the moist W Cape except as vagrant (Friedmann 1948, Irwin 1988, Skead 1995). Intra-African migrant. Arrive with the rains, sing and breed during rainy season, then disappear. In The Gambia and Nigeria they occur in all seasons (Elgood et al. 1994, Barlow et al. 1997). In Liberia there are two records, February and March (Gatter 1997). In Ghana, Togo and southern Benin birds are seen January to August or September, rarely in October and November (Grimes 1987, Cheke and Walsh 1996, Anciaux 2000). In Somalia scarce, records are from May to September (Ash and Miskell 1998). In East Africa they appear with the rains, October to May and

they breed from August and can occur in any month; they are inconspicuous when not in song ( Jackson 1938, Lewis and Pomeroy 1989, Zimmerman et al. 1996, Stevenson and Fanshawe 2002). In Malawi and Zambia they arrive in October and depart by April (Benson and Benson 1977, Aspinwall and Beel 1998). In southern Africa they occur from September to April (Vernon et al. 1997), occasionally over winter (Parker 1994). Fairly common.

Habitat and general habits Open woodland, acacia savanna, avoiding dense evergreen forest and arid regions. They forage through the foliage, and take food on the ground, foraging in cattle dung and hawking termite alates (Eisentraut 1973, Louette 1981, Rowan 1983, Irwin 1988, Skead 1995, RBP). Solitary, shy, not often seen except when males are singing. Males are territorial, sing with and chase other cuckoos (Tarboton 1975, Rowan 1983).

Food Insects, mainly caterpillars, including hairless army worms Spodoptera exempta and other hairy larvae (Rowan 1983, Irwin 1988, Skead 1995).

Breeding During the rains. In West Africa, at Dakar, Senegal, a nestling was seen in July (P. Reynaud, in litt.). In The Gambia birds sing in April and May during onset of rains, where local agriculturists know the bird by its call “dou corr”, the Mandinka word for “work”, a

500 Oriental Cuckoo Cuculus optatus signal to start work in the fields; an adult cuckoo was seen searching for nests at the onset of rains in June and July (Hopkinson 1922, Bannerman 1933, Gore 1990, Barlow et al. 1997, C. R. Barlow, pers. comm.). In coastal Ghana breeding is suggested by song from January to June, in northern Nigeria song and female with oviduct egg from February to April (Bates 1924, Grimes 1987, Elgood et al. 1994), in northern Cameroon birds sing in January (Payne 1986), in Darfur, Sudan, males are in song and female with yolked ovarian follicles found in May and June (Lynes 1925). In Kenya they call from March to June in central region, in Baringo area from August to November and young fledge in August, at Naivasha in September, November and December, in Tsavo bush area young cuckoos fledge by November, in northern Tanzania (Serengeti) they breed in January and February (Lewis and Pomeroy 1989, Zimmerman et al. 1996). In southern Africa the cuckoos sing and breed September to December (Tarboton 1975, Maclean 1993, Skinner 1996,Vernon et al. 1997). Brood-parasitic. Hosts, mainly Fork-tailed Drongo Dicrurus adsimilis (Plowes 1948, Payne and Payne 1967, Tarboton 1975, 1986, Rowan 1983),

also Yellow-billed Shrike Corvinella corvina in West Africa in Senegal (P. Reynaud, in litt., July 1996, Dakar, Senegal), Ghana (Grimes 1979),Togo (Cheke and Walsh 1996) and Nigeria (Elgood et al. 1994), and perhaps other host species rear the young African Cuckoo (Friedmann 1948, 1956, 1967).The eggs are whitish with pale brown spots or rufous blotches, similar to drongo eggs, 25 ⫻ 18 mm (Pitman 1957, Ottow and Duve 1965, Tarboton 1975, 1986, 2001, Rowan 1983). Eggs matching the host eggs are not always accepted, and non-matching eggs in tests are not always removed by the drongo (Tarboton 1986). The incubation period is unknown, somewhere between 11 and 17 days.The nestling cuckoo evicts the host eggs and young. Nestling is in the quill stage by day 8 when the feet are pale yellow, the bill is brown and eyes are open; feathers emerge in 14 days; the bird is well feathered in 19 days and it fledges in 20–23 days (Tarboton 1975). Parasitism rate, 1% of Fork-tailed Drongo nests in southern Africa, locally more often (Payne and Payne 1967, Tarboton 1975). Breeding success was 8% from egg to fledging in one area, 38% in another area (Tarboton 1975, 1986).

Oriental Cuckoo Cuculus optatus Gould, 1845 Cuculus optatus Gould 1845, Proceedings of the Zoological Society of London, 1845, p. 18 (Port Essington, Northern Australia). Monotypic. Other scientific names: Cuculus horsfieldi Moore, 1857. Other common names: Horsfield’s Cuckoo, Dove Cuckoo, Northern Muted Cuckoo.

Description ADULT: Sexes alike (in part), male, above, head and back dark ashy gray, wing dark ashy gray, tail blackish, the central rectrices T1 black with small white spots next to basal half of shaft and a white tip, T2 to T4 with narrow white spots along shaft and white notches on inner web, T5 outer rectrices incompletely barred with white inner notch continuous with large white spot along shaft; underparts, chin to upper breast ashy gray, lower breast

and belly white with black bars 2–4 mm (where nearly as wide as white bars), under tail coverts buff either unbarred or with fine black bars, bend of wing unbarred white, under wing coverts whitish with fine black bars, a gray band formed by the greater wing coverts and a broad whitish band across secondaries and primaries (all but the outer 3 or 4); female, a gray phase similar to adult male, the underparts with ochre tinge on upper breast and usually with narrower ventral black bars; also a rufous phase with upperparts barred chestnut and blackish brown, rump and upper tail coverts with dark bars nearly as broad as the rufous bars, wing barred dark brown and rufous; underparts, barred pale chestnut and blackish, lower breast tinged rufous, the bend of wing white but sometimes obscured by other coverts; eye-ring yellow, iris yellow to brown, bill black, greenish to yellow at base and cutting edge, feet yellow.

Oriental Cuckoo Cuculus optatus 501 JUVENILE: Above, crown and back slate gray with narrow white edges on feathers, nape with white patch, wing coverts slate gray with narrow white edges, wing with brown notches in the primaries and secondaries, tail blackish with terminal white bar and several complete or incomplete white bars across the tail feathers (T1 with white spots along the shaft); underparts, the chin to breast black or black with fine white bars, belly barred black and white, under tail coverts buffy white with incomplete black bars, bend of wing unbarred white, under wing coverts white or nearly all white; either there are two plumage phases or the plumage is variable in both sexes, under wing coverts white or nearly all white; iris brown. NESTLING: Naked at hatching, skin dark brown, the mouth lining orange, with yellow gape edge and a black streak on each side of the palate (Y. Sibnev in Payne 1997,Tojo et al. 2002). SOURCES: AMNH, ANSP, BMNH, BPBM, BYU, CM, DMNH, FMNH, KU, LSU, MCZ, MVZ, MZB, RMNH, ROM, SMF, UMMZ, USNM, UWBM, WAM, YIO, YPM, ZMB, ZMUC, ZRC, ZSM.

History This cuckoo raises the question of whether the large migratory northern form Cuculus optatus is conspecific with the similar southern C. saturatus (Payne 1997b, Higgins 1999). The songs of northern birds are given at a lower pitch than songs of southern birds, the number of notes is less (C. optatus songs are usually a pair of notes) and the song lacks a highpitched note at the beginning, distinct enough to recognize at the species level. Although plumages are similar, sex for sex the northern C. optatus are nearly always larger than the southern C. saturatus. The large cuckoos that migrate from northern and eastern Asia to New Guinea and Australia were known for many years as Cuculus optatus Gould 1845 (Mathews 1918). ANSP has specimens from Gould’s type series, and Stone (in Stone and Mathews 1913) designated ANSP 19980 as the lectotype for Cuculus optatus from the three Gould specimens. The label of ANSP 19980 has “M” and this nota-

tion in addition to the defined locality of Port Essington in coastal Northern Territory may have influenced Stone in selecting a lectotype; no field label indicates that the bird was a male. Junge (1937a) reported a wing of 176.5 mm, as measured by W. Stone at ANSP, and referred C. optatus Gould to the synonymy of C. saturatus Blyth, 1843, the small-winged cuckoo of the Himalayas. Mason (1997) noted this wing measurement was “anomalously small for a male, even for nominotypical saturatus” and rejected Stone’s designation of a lectotype. The measurement of 176.5 mm conflicts with Gould’s “7 3/4 inches” (c. 197 mm) and my measurement of 189 mm of the wing in Gould’s (1845b, plate 84) color plate of “a male of the natural size.” Meyer de Schauensee (1957) earlier had rejected ANSP 19980 as “the type” of optatus because the barred breast agrees “neither with the description or the bird figured,” but in fact Gould described “all the under surface buffy white, crossed by bands of black” and Gould’s color plate shows barred underparts. In the text Gould described the immature as “distinctly barred with black and white.” His color plate shows an adult with two others in the background and is idealized with all three birds in adult plumage, even though his description mentions the juvenile plumage as well as the adult. The right wing of ANSP 19980 measures 176 mm but lacks the longest primaries (P7 and P8).The left wing measures 190 with the tip (perhaps 2 mm) of the longest primary missing; the original length is about 191, close to Gould’s 189 mm. The specimen is in 65% adult plumage and 35% juvenile plumage (not “juvenile plumage” as stated on a later label, Mason 1997). Two other ANSP specimens, 20007 from “Australia” and 19981 from “New South Wales” are from Gould’s type series. The first is an adult with wing of 194 (right) or 196 (left), and the second is in mixed plumage (half juvenile, half adult) with wing 206 mm (right) or 202 (left). ANSP 19980, the lectotype of Cuculus optatus Gould 1845, is large enough to be the northern migratory cuckoo which on its breeding grounds in northern Asia have a wing of this size; and this name has priority over Cuculus horsfieldi Moore 1857.

502 Oriental Cuckoo Cuculus optatus These large cuckoos have sometimes been recognized as Cuculus canoroides S. Müller 1845. Müller did not designate a holotype, and it was uncertain whether the name applied to the large or small migratory cuckoos; he measured the five specimens in his type series as wings 190-214. From this series Junge (1956) designated a specimen (Schlegel catalogue s.n. Cuculus striatus no. 34, RMNH 88213) taken on Gulu Doesoen near west Borneo in August 1836, wing 187 mm, as the lectotype of canoroides, and he determined that 88213 was similar to males of the small Asian saturatus ( Junge 1937a). He described the specimen as an immature brown- (“red”) phased male as lectotype; by this act the name canoroides became a synonym of Cuculus saturatus. Mathews (1912, 1918) summarized other nomenclatural comments. Designation of lectotypes of optatus by Stone (1913) and canoroides by Junge (1956), together with examination of the ANSP specimens, clarifies the priority in name of these large migratory cuckoos. In their breeding ranges, adult C. optatus are larger (M, wing 190–227) than adult C. saturatus (M, wing 174–194). In their wintering ranges, nearly all birds can be distinguished by size. Passage and wintering birds with retained juvenile feathers can be distinguished by plumage (C. optatus, juvenile head, throat and breast nearly black with fine whitish edge; C. saturatus, juvenile black with broad white bars). C. optatus were regarded as indistinguishable from C. saturatus by White and Bruce (1986) on the basis of birds of intermediate size in India, Russia and China (Wells 1972), and Vaurie (1965) stated that the two intergrade in China, but these sources gave no data.The comments were based in part on misidentified specimens (AMNH 624917, FMNH 229662, USNM 296269, 296261, 269266 and 269263) that are in fact C. canorus, which average intermediate in size between C. saturatus and C. optatus. Nevertheless, there is some overlap in wing lengths of these two cuckoos.

Measurements and weights Japan, Okinawa and Kurile Is, wing, M (n ⫽ 20) 194–222 (206.2 ⫾ 8.2), F (n ⫽ 22) 185–211 (196.2 ⫾ 8.8); tail, M 126–181 (150.0 ⫾ 15.2), F

130–171 (147.3 ⫾ 15.8); bill, M (n ⫽ 17) 19.7– 24.6 (22.0 ⫾ 1.7), F (n ⫽ 18) 17.9–23.7 (20.9 ⫾ 1.7); tarsus, M (n ⫽ 9) 18.2–22.4 (20.8 ⫾ 1.2), F (n ⫽ 10) 17.1–24.2 (20.2 ⫾ 2.0) (5 MM had wing ⬍ 200, 6 FF had wing ⬍ 190) (BMNH, FMNH, MCZ, MVZ, ROM, USNM,YIO); Russia and Kazakhstan, wing, M (n ⫽ 16) 190–213 (204.8 ⫾ 2.7), F (n ⫽ 14) 184–211 (195.1 ⫾ 8.2); tail, M 128–173 (156.7 ⫾ 11.8), F 128–168 (142.6 ⫾ 11.8) (AMNH, BMNH, MCZ, UMMZ, UWBM, ZMUC, ZSM); N China (Sichuan and continental NE China north of 30o N), wing, M (n ⫽ 30) (185) 194–227 (206.4 ⫾ 8.2), F (n ⫽ 16) 185–204 (194.8 ⫾ 8.7); tail, M 132–171 (159.6 ⫾ 17.9), F 127–169 (153.2 ⫾ 14.1) (M wing 185, Chefoo, 6.6.1873, BMNH 87.12.19.26, included; M wing 180, Shandong, CM 40888, excluded; all other males wing ⱖ 190; largest males, wing 225–227) (AMNH, BMNH, CAS, CM, FMNH, MCZ, USNM, ZMUC); Weights, Japan, U 70.9-107.3 (101.2) (Enomoto 1941); Korea, M (n ⫽ 2) 108–122 (115), F (n ⫽ 1) 119 (MVZ); the Philippines, M (n ⫽ 13) 76.4– 118.1 (99.0), F (n ⫽ 11) 71.2–136.4 (89.1) (BPBM, CMNH, FMNH, USNM,YPM). Wing formula, P8 > 9 > 7 > 6 > 5 > 4 > 10 > 3 > 2 > 1.

Field characters Overall length 32-33 cm. Plumage above gray, belly and under tail coverts buff (not white as in Common Cuckoo C. canorus), and in central and eastern Asia where the two species occur together the upperparts are darker gray and the lower breast and belly have black bars wider than in C. canorus. Bend of wing unbarred white, under wing coverts white or (secondary coverts) whitish with fine black bars (under wing coverts more heavily barred in C. canorus), the white band broad from inner secondaries through P1-7. Rufous phase has broad dark bars on rump (broad rufous bars in C. canorus). Adults are distinguishable by size from the smaller C. saturatus. Juvenile C. optatus and C. saturatus differ in plumage, C. optatus being darker on throat and breast: C. saturatus have more conspicuous white feather tips on crown, back, and wing coverts, and

Oriental Cuckoo Cuculus optatus 503 tips of inner secondaries (1 mm) than C. optatus (0.5 mm). Juveniles are blackish on breast, black bars on lower breast and belly are broader than in juvenile C. canorus, and have the under wing pattern of the adults; juveniles differ in color from the paler juvenile C. canorus, which have brown bars on the underparts. Nestlings differ from other Asian cuckoos in having a black streak on the red palate. Song is soft or muted and lower in pitch than songs of C. canorus; name in Russia is “Muted Cuckoo” or “Dove Cuckoo”. Song differs from that of C. saturatus by being a low-pitched two-note affair and without a higher introductory note.

Voice Song, whistled paired notes “hoop, hoop” in Japan, Korea, Finland, western Siberia, the Yenisei River, Novosibirsk, Lake Baikal, Amur and Ussuria in Russia, Mongolia and Heilongjiang in China).The paired “hoop” notes are 0.1 sec long at a pitch of c. 430 Hz, lower than in songs of C. s. saturatus in southern Asia. (Seebohm 1882, Johansen 1955, Mauersberger 1980, Schubert 1982, Cramp 1985, Ernst 1992a,b, 1996, Martens and Eck 1995, Hwang 1997, P. Alström recordings). Songs in Finland sometimes have paired notes with the first higher in pitch (450 Hz) than the second (420 Hz). Other calls include hoarse croaks, chuckles and), a harsh “gaak-gaak-gak-ak-ak-ak” and a bubbling female “quick-quick-quick”(Cramp 1985, Lindholm and Lindén 2003).

Range and status Migrant C. optatus occur in summer from Finland and Komi west of the Ural Mts in Russia north and east to the arctic circle on the Lena River, east to the Pacific coast on the Okhotsk Sea, the Kurile Islands and Sakhalin, south to Kazakhstan, the Altai Mts, Mongolia, northern China, Korea and Japan (Yamashina 1931, Vaurie 1965, Gore and Won 1971, Cramp 1985, Tomek 1985, Cheng 1991, Brazil 1991, Fiebig 1993, Hagemeijer and Blair 1997,Yoon 2000; AMNH, BMNH, SMF, UWBM, ZMUC, ZSM). Birds taken in June on Ishigakijima, between Taiwan and Okinawa (AMNH 625178, 625179, wing, M 195, F 185; Ogawa 1905), indicate

breeding C. optatus in the Nansei-shoto islands. In Russia, distributions of birds throughout the Urals and Pechora River region west to the Vychegda River, and Russian records are mapped by Hagemeijer and Blair (1997) and supported with specimens and tape recordings (Grote 1927, Danilowitsch 1934, ZMM), and in Finland with photographs and tape recordings (Vasamies 1998, Topp 1999, Lindholm and Lindén 2003). Cuckoos arrive on the breeding grounds in Japan in late April (Yoshino 1999). At end of the season they move southward through Okinawa and elsewhere in the Nansei-shoto (Ryu Kyu) islands, Taiwan and Botel Tobago and southeast Asia to the wintering grounds (Hachisuka and Udagawa 1951, Severinghaus and Blackshaw 1976, McWhirter et al. 1996). Mainly passage migrants in southeast Asia and Malay Peninsula, some birds winter in Vietnam, the Philippines and Micronesia (Palau I, Yap I). Migrant or winter south to Greater Sundas (Sumatra, Nias, Borneo),Wallacea from Sulawesi to the Lesser Sundas (Flores, Sumba, Sumbawa) and Timor, New Guinea, New Britain, Lord Howe I, Solomon Is and N and E Australia (Rensch 1931, Stresemann 1940, Mayr 1944b, Baker 1951, Galbraith and Galbraith 1962, Gilliard and LeCroy 1967a, Diamond 1972, Pratt et al. 1977, Blakers et al. 1984, Coates 1985, Cramp 1985, White and Bruce 1986, van Marle and Voous 1988, Doughty et al. 1999, Higgins 1999, Kennedy et al. 2000; AMNH, BMNH, MNHN, MZB, RMNH, SMF, ZRC). In Indonesia, the proportion of C. optatus to C. saturatus in winter increases from west to east, with 17% C. optatus in Sumatra, 29-46% in Java, c. 50% in Borneo, 92% in Sulawesi and Lesser Sundas, 88% in the northern Moluccas, and 100% in the southern Moluccas and Timor (Cramp 1985). In the Philippines C. optatus are known from more specimens and more islands than are C. saturatus (Dickinson et al. 1991, Miranda et al. 1999). In New Guinea most records are in coastal and river lowlands, birds are also in eastern highlands over 1800 m (Mayr and Rand 1937, Rand 1942a,b, Mayr and Gilliard 1954, Gilliard and LeCroy 1966, 1967b, Diamond 1972, BMNH); a small male taken in the D’Entrecasteaux Is in March had wing 188 (USNM 584876) and might be C. saturatus,

504 Oriental Cuckoo Cuculus optatus

however in their breeding range C. optatus are as small as wing 184. In Australia C. optatus occur through the northern winter from November to May (Blakers et al. 1984), near Darwin from December to early April (Crawford 1972), in the Kimberley Region from October to April with a few birds through June to August (Storr 1980, Johnstone and Burbidge 1991), and rarely in South Australia (Read 1995). In New Zealand they are uncommon, with most records from October to March (Heather and Robertson 1997, Higgins 1999). Errant passage migrants turn up in the Bering Sea islands and western Alaska (AOU 1998; USNM); one appeared in Israel in August (Shirihai 1996). Breeding densities, in western Russia in pine Pinus and spruce Picea forests, c. 0.4–0.5 males per km2, in spruce-aspen forests of lower Vychegda region 0.1 males per km2; estimated European population 5000–10,000 breeding pairs (Hagemeijer and Blair 1997).

Habitat and general habits Eurasian coniferous and mixed coniferous-deciduous forest, birch and aspen, larch taiga, and broad-leafed

forests and thickets. In Russia and eastern Siberia they are more a forest bird than Common Cuckoo C. canorus and in the southern Urals they occur in small woodlands and shrubs (Cramp 1985, Ernst 1996, Hagemeijer and Blair 1997). In Sichuan they occur to 4500 m, in Altai at lower altitudes than C. canorus, rarely above 1000 m (Dement’ev and Gladkov 1966, Cramp 1985). In Japan they are in broad-leafed forest or mixed forest (Brazil 1991). Wintering habitat in the Malay Archipelago includes woodlands near streams in hill country, in Wallacea and New Guinea the habitat is not well known because of confusion of cuckoo species (White and Bruce 1986, Coates and Bishop 1997). In Australia, wintering, birds in open woodland are usually alone or in twos, sometimes in groups near a patch of food (Bell 1979, Storr 1980, Blakers et al. 1984).They feed on branches and in foliage in trees and take caterpillars from clumps of mistletoe. They capture cicadas emerging on trunks and large branches of trees by flying from a perch to seize the insect. They also perch on a branch and fly to the ground to catch insects, sally to other trees and to grass under the trees, spot a caterpillar as far as 50 m (Bravery 1967, 1970, Pizzey and Knight 1998, Higgins 1999).

Himalayan Cuckoo Cuculus saturatus 505

Food Insects, including large stick insects, large and hairy caterpillars, grasshoppers, beetles, flies, wasps, and ants (Dement’ev and Gladkov 1966, Cramp 1985).

Breeding During spring when their warbler hosts breed, in central Russia in June and July (Dement’ev and Gladkov 1966, Cramp 1985), and in Japan from early May to late June (Yoshino 1999). Brood-parasitic. Hosts in Russia include Chiffchaff Phylloscopus collybita, Eastern Crowned-warbler Phylloscopus (reguloides) coronatus, Pale-legged Leaf-Warbler P. tellennipes and other Phylloscopus species, Tree Pipit Anthus hodgsoni and Black-

throated Accentor Prunella atrogularis (Irisov 1967, Cramp 1985, Knystautas and Sibnev 1987, Sibnev in Payne 1997). In Japan, cuckoos are reared by Bush Warbler Cettia diphone in Hokkaido, and by crowned-warbler, Narcissus Flycatcher Ficedula narcissina and Short-tailed Bush Warblers (Asian Stubtail) Cettia (Urosphena) squameiceps in Honshu (Higuchi 1998). Eggs are reddish brown in Hokkaido where Bush Warblers have similar brown eggs; white or bluish white with brown markings elsewhere in Japan (Makatsch 1971, Higuchi and Sato 1984, Higuchi 1998,Yoshino 1999,Tojo et al. 2002). Eggs are 22.1 ⫻ 15.8 mm in Hokkaido, 19.5 ⫻ 14.4 mm in Honshu (Higuchi 1998). The incubation and nestling periods are unknown.

Himalayan Cuckoo Cuculus saturatus Blyth, 1843 Cuculus saturatus Blyth, 1843, Journal of the Asiatic Society of Bengal, 12, p. 942. (Nepal) Monotypic. Other common names: Blyth’s Cuckoo, Mountain Cuckoo.

the rufous bars, wing barred dark brown and rufous; underparts, barred pale buff and blackish, lower breast tinged rufous, the bend of wing white (sometimes obscured by other coverts); eye-ring yellow, iris yellow to brown, bill black, greenish to yellow at base and cutting edge, feet yellow.

Description ADULT: Sexes alike (in part), male, above, head and back dark ashy gray, wing dark ashy gray, tail blackish, the central rectrices T1 black with small white spots next to basal half of shaft and a white tip, T2 to T4 with narrow white spots along shaft and white notches on inner web, T5 outer rectrices incompletely barred with white inner notch continuous with large white spot along shaft; underparts, chin to upper breast ashy gray, lower breast and belly white with black bars 2–4 mm (nearly as wide as white bars), under tail coverts buff either unbarred or with fine black bars, bend of wing unbarred white, under wing coverts whitish with fine black bars, a gray band formed by the greater wing coverts and a broad whitish band across secondaries and primaries (all but the outer 3 or 4); female, a gray phase similar to adult male, the underparts with ochre tinge on upper breast and usually with narrower black ventral bars; also a rufous phase with upperparts barred chestnut and blackish brown, rump and upper tail coverts with dark bars nearly as broad as

JUVENILE: Above, crown and back slate gray with narrow white edges on feathers, nape with white patch, wing coverts slate gray with narrow white edges, wing with brown notches in the primaries and secondaries, tail blackish with terminal white bar and several complete or incomplete white bars across the tail feathers (T1 with white spots along the shaft); underparts, the chin to breast black or black with fine white bars, belly barred black and white, under tail coverts buffy white with incomplete black bars, bend of wing unbarred white, under wing coverts white or nearly all white; either there are two plumage phases or the plumage is variable in both sexes, under wing coverts white or nearly all white; iris brown. NESTLING: Not described. The nestlings described in the Malay Peninsula, Java and Borneo (Madoc 1956b, Becking 1975b, 1981, Moyle et al. 2001) are C. lepidus; not C. saturatus, which does not breed in those areas.

506 Himalayan Cuckoo Cuculus saturatus SOURCES: AMNH, ANSP, BMNH, CM, CMNH, FMNH, KU, LSU, MCZ, MZB, NSM, RMNH, ROM, SMF, TISTR, UMMZ, USNM, UWBM,YPM, ZMA, ZMB, ZRC, ZSM.

Measurements and weights N India (Kashmir, Himachal Pradesh, Uttar Pradesh, Bihar, Assam) and Nepal, wing, M (n ⫽ 35) 174–194 (183.4 ⫾ 5.8), F (n ⫽ 15) 163–190 (175.4 ⫾ 8.2); tail, M 113–152 (136.4 ⫾ 11.7), F 113–152 (135.6 ⫾ 10.6); bill, M (n ⫽ 21) 17.9–22.1 (20.5 ⫾ 1.1), F (n ⫽ 11) 18.5–23.0 (20.0 ⫾ 1.4); tarsus, M 14.2–20.2 (17.2 ⫾ 2.3), F 14.2–20.4 (17.2) (birds with wing longer than 188 (n ⫽ 5) (191–194) were taken March to May and perhaps included migrant Oriental Cuckoo C. optatus) (AMNH, FMNH, LSU, ROM, UMMZ,YPM, ZSM); Taiwan, Hainan and southern coastal China south of 26° N (Fujien, Guangdong, Hong Kong), wing, M (n ⫽ 20) 172–195 (185.6 ⫾ 5.5), F (n ⫽ 19) 168–194 (179.0 ⫾ 8.0), U (n ⫽ 1) 180 (includes April and May, perhaps some migrant C. optatus; excludes F on 5 May with wing 199); 18 May to 10 September, M (n ⫽ 4) 172–191 (182.2 ⫾ 7.7), F (n ⫽ 9) 167–190 (180.9 ⫾ 9.5) (AMNH, BMNH, CAS, MCZ, USNM,YPM). Weights: Taiwan, M (n ⫽ 9) 75–103 (93.1), F (n ⫽ 7) 60–80 (80.4) (Cramp 1985), India and Nepal, M (n ⫽ 3) 96–105 (99.8), F (n ⫽ 2) 72–77.8 (74.9) (AMNH, KU, ZSM), the Philippines, M (n ⫽ 3) 73.2–82.2 (78.7), F (n ⫽ 5) 65–87 (72.6) (CMNH). Wing formula: P8 ⬎ 9 ⱖ7 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 10 ⬎ 3 ⬎ 2 ⬎ 1.

(1 mm) than C. optatus (0.5 mm); and juveniles have black bars on lower breast and belly broader than in juvenile C. canorus. C. lepidus are smaller and darker buffy below. Best distinguished from other cuckoos by song, often with 3 or 4 “hoop” notes, higher in pitch than C. optatus and lower than C. lepidus and has a high soft introductory note; the “hoop” notes are on a single pitch, unlike song of C. canorus.

Voice A whistled note followed by 2, 3 or 4 lower mellow notes “hoop, hoop-hoop” in Pakistan, Kashmir, India, Nepal, Burma and eastern China (Fujian) ( Jerdon 1862, Stevens 1925, La Touche 1931–34, Stresemann and Heinrich 1939, Baker in Bent 1940, Frome 1946, Bates and Lowther 1952, Smythies 1949, 1953, King and Dickinson 1975, Wells and Becking 1975, Severinghaus and Blackshaw 1976, White 1984, Roberts 1991, Connop 1993, Martens and Eck 1995, King 2005, P.Alström recordings, RBP recordings). The introductory note is short, slightly higher and not so loud, c. 480 Hz in India and Nepal; the other notes are 0.1 sec long. In central China the songs in Sichuan and Shaanxi are 3- or 4-note affairs and have a higherpitched introductory note, as in birds of the Himalayas. Other calls include hoarse croaks, chuckles, a harsh “gaak-gaak-gak-ak-ak-ak” and a bubbling female “quick-quick-quick”. Birds give songs of 2, 3, and 4 “hoop” notes, with the number of notes sometimes varying within a single male’s song.

Range and status Field characters Overall length 32–33 cm. Plumage above gray, belly and under tail coverts buff (not white as in Common Cuckoo C. canorus; not dark buff as in Sunda Lesser Cuckoo C. lepidus). Bend of wing unbarred white, under wing coverts white or (secondary coverts) whitish with fine black bars (under wing coverts more heavily barred in C. canorus), under wing with broad white band from inner secondaries through P1-7. Rufous phase has broad dark bars on rump (broad rufous bars in C. canorus). Juveniles have more conspicuous white tips on crown, back, and wing coverts, and tips of inner secondaries

Breed in northern Pakistan (Murree Hills) through the Himalayas east to Nepal and Arunachal Pradesh in India, southeast Asia, southern China (Fujian, Guangxi), Taiwan and Hainan (Vaurie 1965, Ali and Ripley 1969, Abdulali 1977, Guangdong et al. 1983, Cramp 1985, Cheng 1991, Lekagul and Round 1991, Roberts 1991, Grimmett et al. 1999; AMNH, BMNH, RMNH). Summer residents occur in Kashmir from late April to August (Baker, in Bent 1940), in Nepal they occur from March to September, common in Himalaya (Inskipp and Inskipp 1985, Martens and Eck 1995), and in Assam (Khasi Hills, Garo Hills, Naga

Himalayan Cuckoo Cuculus saturatus 507

Hills) they occur March to September (UMMZ). They are partial migrants in southern Asia, where birds are observed in the breeding season and less often in winter in the plains of northern India; migrants in spring appear later at the higher elevations (Stevens 1925). Common and widespread in the Himalayas. In Burma, they occur in the Chin Hills and the northeast (Smythies 1940, 1953). In China, they occur in the breeding season on Hainan, on the mainland south of 26° N latitude. In the southern coastal plains both C. saturatus and C. optatus were taken in May when the latter were in migration northwards, and in September when they were moving southwards. The scarcity of known breeding birds, the overlap of size of the two species, and the scarcity of song recordings make it uncertain whether both species occur together in central China in the breeding season. In Yunnan both C. canorus and C. saturatus occur in the breeding season, in the LiKiang mountains and source of the Salween and Mekong rivers (AMNH, USNM). In Taiwan they are most common on Mt Ari, as summer visitors from April to September (Swinhoe 1863, Hachisuka and Udagawa 1951). Swinhoe (1863), resident naturalist on Taiwan, collected a specimen there and named it Cuculus kelungensis. He noted that birds on the mainland near Amoy, C. canorus, had a song unlike the birds of Taiwan and the Himalayas, C. saturatus. David and Oustalet (1877) noted that these latter birds had a call of 3 or 4 notes “hou-hou-hou-hou” differing from C. canorus.

In China, they occur in the breeding season in coastal area south of 26o N. Both C. saturatus and the larger C. optatus were taken in May when large birds were in migration northwards, and in September when large birds were moving southwards. The scarcity of known breeding birds and song recordings make it uncertain whether these two cuckoo species occur together in the breeding season; they are not known to do so. In migration, C. saturatus occur in southeast Asia,Thailand and the Malay Peninsula and Singapore. In winter they occur south of the continent in the Andaman Is, Sumatra (Utara, Barat, Malacca Straits, Simeule, Nias), Java (mainly western, most records on Gede Halimun and Gede Pangrango), Bali, Borneo (Labuan I, Sabah), and the southern Philippines (Mindanao, Mindoro, Negros, Nangsi; rare migrant on Luzon at Dalton Pass) (Becking 1975b, van Marle and Voous 1988, Dickinson et al. 1999, Wells 1999, Sheldon et al. 2001, Robson 2002; UWBM), and there are a few winter records in northern and peninsular India, northern Sulawesi, Halmahera, and western New Guinea (Waigeo, Vogelkop Peninsula). Other records in southern New Guinea and in Australia of Cuculus (wing 183192) are this species or more likely are small C. optatus. Breeding density near Mussoorie, India, 3 calling birds per mile2 (c. 1 bird per km2) (Fleming 1967).

Habitat and general habits Forest, in the Himalayas they are in mixed coniferous-deciduous forest (birch and aspen), larch taiga, and in broad-leafed forests and thickets, oak rhododendron forests and wooded areas, and in subtropical woodlands near streams in hill country. In northern montane Pakistan they are in mixed deciduous and coniferous forest above 1800 m (not in the alpine region where Common Cuckoo Cuculus canorus occur, nor in subtropical pine where Indian Cuckoo C. micropterus and Gray-bellied Cuckoo Cacomantis passerinus occur) (Roberts 1991), in Kashmir in birch above tree-line (Price and Jamdar 1990), in Nepal in coniferous forest mainly 1500–3300 m (Inskipp and Inskipp 1985) and 1830–3000 m (Martens and Eck 1995), in Guangxi, China, at 1400–1600 m in Fagus beech

508 Sunda Lesser Cuckoo Cuculus lepidus forest (Vuilleumier 1993) and in mixed pine and oak in the Chin Hills, Burma (Robson et al. 1998). In the Malay Peninsula, migrant C. saturatus are in coastal and plains forests; resident C. lepidus live in mid and canopy strata of tall montane forest (Wells 1999). In Java and Sumatra, they winter in lowland rain forest, secondary forest and cultivation, and offshore islands (Becking 1975b). In Wallacea, wintering cuckoos occur in lowland forest, forest edge, woodland, lightly wooded cultivation and savanna (Rensch 1931, Coates and Bishop 1997).

Food Insects, mainly caterpillars (hairless and hairy), grasshoppers, crickets, cicadas, beetles, flies and ants; fruits, and the shoots and needles of pines (Stevens 1925).

Breeding Season coincides with that of their warbler hosts. In India, C. saturatus breed mainly in May and June (Baker 1934). In Nepal, they breed from March to August (Inskipp and Inskipp 1985), in China (Fujian) in May (La Touche 1931–34), and they sing in the Chin Hills, Burma, in April (Robson et al. 1998). Brood-parasitic. Hosts, small warblers of the genera Seicercus and Phylloscopus, including Western Crowned-warbler Phylloscopus occipitalis and Blyth’s Leaf-Warbler P. reguloides (Baker 1906, 1927, 1934, Bates and Lowther 1952, Ali and Ripley 1969, Becking 1981). Eggs are pale blue or white with fine black and brown or reddish stippling, known from oviduct eggs and similar eggs in host nests, 21.1 ⫻ 15.6 mm (Baker 1906, 1927, 1934). The incubation and nestling periods are unknown.

Sunda Lesser Cuckoo Cuculus lepidus S. Müller, 1845 Cuculus lepidus S. Müller, 1845, Verhandelingen over de Natuurlijke Geschiedenis der Nederlandsche overzeesche bezittingen, Land-en Volkenkunde, 8, p. 236. Polytypic.Two subspecies. Cuculus lepidus lepidus S. Müller, 1845, Cuculus lepidus insulindae Hartert, 1912. Other common names: Sunda Cuckoo, Himalayan Cuckoo (part). Othr names: Cuculus saturatus lepidus.

Description

phase similar to adult male, the underparts with ochre tinge on upper breast and usually with narrower ventral black bars; also a rufous phase with upperparts barred chestnut and blackish brown, rump and upper tail coverts with dark bars nearly as broad as the rufous bars, wing barred dark brown and rufous; underparts, barred pale chestnut and blackish, lower breast tinged rufous, the bend of wing white; eyering yellow, iris yellow to brown, bill black, greenish to yellow at base and cutting edge, feet yellow.

ADULT: Sexes alike (in part), male, above, head and back dark slaty gray, wing dark ashy gray, tail blackish, the central rectrices T1 black with small white spots next to basal half of shaft and a white tip, T2 to T4 with narrow white spots along shaft and white notches on inner web, T5 outer rectrices incompletely barred with white inner notch continuous with large white spot along shaft; underparts, chin to upper breast ashy gray, lower breast and belly buff to light rufous with black bars 2–4 mm (nearly as wide as white bars), under tail coverts dark buff or rufous either unbarred or with fine black bars, bend of wing unbarred white, under wing coverts whitish with fine black bars, a gray band formed by the greater wing coverts and a broad whitish band across secondaries and primaries (all but the outer 3 or 4); female, a gray

JUVENILE: above, crown and back slate gray with narrow white edges on feathers, nape with white patch, wing coverts slate gray with narrow white edges, wing with brown notches in the primaries and secondaries, tail blackish with terminal white bar and several complete or incomplete white bars across the tail feathers (T1 with white spots along the shaft); underparts, the chin to breast black or black with fine white bars, belly barred black and white, under tail coverts buffy white with incomplete black bars, bend of wing unbarred white, under wing coverts white or nearly all white; either there are two plumage phases or the plumage is variable in both sexes, under wing coverts white or nearly all white; iris brown.

Sunda Lesser Cuckoo Cuculus lepidus 509 NESTLING: Naked at hatching, skin dark brown, the mouth lining orange, inner edge of gape bright red SOURCES: AMNH, ANSP, BMNH, MZB, NSM, RMNH, SMF, USNM,WAM, ZMA, ZRC, ZSM.

Subspecies Cuculus lepidus lepidus S. Müller, 1845; plumage whitish below with dark buffy under tail coverts; Malay Peninsula, Sumatra, Java, and the Lesser Sunda Islands eastward to Timor. Cuculus lepidus insulindae Hartert, 1912; plumage darker below with dark buffy underparts and buff to rufous on lower belly and under tail coverts; Borneo.

Measurements C. l. lepidus: Malay Peninsula, Sumatra, Java and Bali, wing, M (n ⫽ 16) 141–166 (154.1 ⫾ 6.7), F (n ⫽ 20) 137–155 (144.8⫾5.2); tail, M 114–134 (122.2 ⫾ 7.8), F 103–117 (113.2 ⫾ 10.3) (excludes wintering migrant birds with wing ⬎ 164 taken October to March) (AMNH, RMNH, SMF, ZSM); C. l. insulindae, Borneo, wing, M (n ⫽ 6) 145– 152 (147.2 ⫾ 2.6), F (n ⫽ 4) 134–152 (142.8 ⫾ 7.2); tail, M 146–152 (147.2 ⫾ 2.6), F 128–164 (141.3) (AMNH, ZRC). Wing formula: P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 10 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 29–30 cm. Plumage above slaty gray, belly and under tail coverts rich buff (not pale as in Himalayan Cuckoo C. saturatus), little contrast between sides of neck and dark upper wing-coverts (more contrast in C. saturatus, where neck is paler). Bend of wing unbarred white, under wing coverts white or (secondary coverts) whitish with fine black bars, a broad white band from inner secondaries through P1-7. Rufous phase has broad dark bars on rump. Juveniles have broad black bars on lower breast and belly. Song similar to C. saturatus but higher in pitch.

Voice A whistled note followed by 2 or 3 lower mellow notes “hoop, hoop-hoop” (Madoc 1956a, Wells

1982, 1999, Wells and Becking 1975, Moyle et al. 2001, Sheldon et al. 2001, King 2005). Although songs of C. lepidus have the same pattern as songs of C. saturatus, songs of C. lepidus differ in their higher pitch, both a high introductory note, and the “hoop” notes which are c. 600 Hz, perhaps due to the smaller body size.To experienced ornithologists the songs sound different in the field (B. King, P. Round).

Range and status Resident in the Malay Peninsula, Sumatra (van Marle and Voous 1988), Java, Bali,), the Lesser Sundas (Lombok, Sumbawa, Sumba, Flores and Pantar) and Timor), Borneo, and in the Moluccas only on Seram (a short-tailed, short-winged juvenile (MZB 2624, taken by Vorderman in 1897) (Vorderman 1898, Rensch 1931, White and Bruce 1986, van Marle and Voous 1988, Smythies 1999,Wells 1999, Sheldon et al. 2001, King 2005; AMNH, ANSP, BMNH, MZB, NSM, RMNH, SMF, USNM, WAM, ZMA, ZRC).

Habitat and general habits Forest. In the Malay Peninsula, C. lepidus live in mid and canopy strata of tall montane forest at altitudes ranging 950–1700 m (Wells 1999). In Sumatra and Java they are in montane forest above 1000 m (Becking 1975b, Prawiradilaga et al. 2003). In Borneo, they are common in the mountains, on Mt Kinabalu from 1200-2750 m (Smythies 1999). In Wallacea, they are in montane forest above 1000 m (Rensch 1931, Coates and Bishop 1997).

Food Insects, mainly caterpillars (hairless and hairy), grasshoppers, crickets, cicadas, beetles, flies and ants; and fruits (Coates 1985, Sody 1989).

Breeding Season coincides with that of their warbler hosts. In the Malay Peninsula, adult cuckoos call from January to July, eggs occur February to June and fledglings appear from late February to late July (Madoc 1956a,b, Wells 1999). In Java, they lay in February, April to July, September and October (Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967), in

510 Common Cuckoo Cuculus canorus

North Borneo on Mt Kinabalu, they sing in February and March (Smythies 1999), eggs appear in April (Moyle et al. 2001) and birds may also breed in July (fledgling perhaps this cuckoo seen in August, Smythies 1999).

Brood-parasitic. Hosts, in the Malay Peninsula they use Chestnut-crowned Warbler Seicercus castaniceps (Madoc 1956a, b,Wells 1999). Most identifications have been based on the eggs illustrated in Hoogerwerf (1949); a photograph (Madoc 1956b) of a young cuckoo reared by Chestnut-crowned Warbler is this species. In Java, hosts include Sunda Bush-Warbler Cettia vulcania, Sunda Warbler Seicerecus grammiceps and Mountain Leaf-Warbler Phylloscopus trivirgatus (Hoogerwerf 1949, Hellebrekers and Hoogerwerf 1967, Becking 1975b, 1981). In Sabah, hosts are P. trivirgatus and Yellow-breasted Warbler Seicercus montis (Moyle et al. 2001). Another record of P. trivirgatus feeding “a young cuckoo speckled black and white” (Phillipps 1970) appears by description to be this cuckoo species. Eggs are creamy white with fine brown spots, 21.2 ⫻ 13.7 mm (Hoogerwerf 1949) and 19 ⫻ 11.8 mm (Madoc 1956a, b).The incubation and nestling periods are unknown.

Common Cuckoo Cuculus canorus Linnaeus, 1758 Cuculus canorus Linnaeus, 1758, Systema Naturae (ed. 10), 1, 10. (Sweden) Other common names: Cuckoo, Grey Cuckoo, Eurasian Cuckoo, European Cuckoo Polytypic. Four subspecies. Cuculus canorus canorus Linnaeus, 1758; Cuculus canorus bakeri Hartert, 1912; Cuculus canorus subtelephonus Zarudny, 1914; Cuculus canorus bangsi Oberholser, 1919.

Description ADULT: Sexes similar (in part); male, above, dark ashy gray, wing ashy gray, tail blackish brown, spotted and tipped with white, no subterminal black band and unevenly barred black, central rectrices T1 and T2 unmarked,T3 to T5 black in center with white spots along the feather shaft, T5 gray with white spots along both edges, sometimes with complete white bar, the tail appearing barred from below; underparts, chin to breast ashy gray, belly and under tail coverts white with black bars, bend of wing white with dark gray bars, under wing coverts barred dark gray and white, a gray band formed by the greater wing coverts, and a broad white stripe across base of

secondaries and a narrow stripe across base of inner primaries (extending to P3 or 4); eye-ring yellow, iris light brown to orange, bill black to greenish black, sometimes with a yellow green base, feet yellow. Female, two color phases occur, a gray phase similar to the adult male, the underparts with ochre tinge on upper breast and usually with narrower ventral black bars; and a rufous (“hepatic”) phase with upperparts barred chestnut and blackish brown, rump and upper tail coverts light rufous or light rufous with blackish brown spots or broken incomplete bars, the rufous bars broader than the dark bars, wing barred dark brown and rufous; underparts, barred pale chestnut and blackish, lower breast tinged rufous. Rufous phase is more numerous in Central Europe than elsewhere (Voipio 1953).A few females in Russia are intermediate in plumage between the gray and rufous color phases (ZMM). JUVENILE: Above crown gray with white bars or tips of feathers, white nape spot, back barred gray and brown with white tips to feathers on upper back, rump barred gray and brown, wing barred gray and

Common Cuckoo Cuculus canorus 511 brown, tail with brown bars and small white oblique, elongate spots along shaft of tail feathers, underparts, throat whitish with blackish bars, breast to under tail coverts whitish with thin black bars, tail appears more barred below than in the adult, some juveniles with barred plumage like that of rufous-phase female but with white edges to feathers on upperparts and with plumage intermediate in barring between gray and rufous adults; iris brown. NESTLING: Naked at hatching, skin flesh-color becoming blackish after three days; gape orange changing to orange-red and gape flanges yellow to orange. SOURCES: AMNH, ANSP, BMNH, BWYO, CM, FMNH, LACM, MCZ, MVZ, NMK, ROM, SMTD, UMMZ, USNM, UWBM, YPM, ZMM, ZSM, ZMB, ZMUC.

History and subspecies Common Cuckoos have been known as broodparasites from classical times. Their brood-parasitic behavior was described in detail by Jenner (1788). The species name refers to the song, the Latin canorus, melodious, or canere, to sing. Cuculus canorus canorus Linnaeus, 1758; plumage above dark ashy gray, underparts white with black bars (2.0–2.5 mm) nearly as wide as white bars; Europe from British Isles, Scandinavia and northern Russia, south to the Pyrenees, Mediterranean, Asia Minor, Asia from Ural Mts east to Siberia, Japan, Korea and northern China; Cuculus canorus subtelephonus Zarudny, 1914; plumage paler above than in C. c. canorus, underparts white with narrow black bars (0.7–1.8 mm); central Asia from Iran, Afghanistan, Baluchistan and Turkmenistan to Xinjiang and A-la Shan in northern China. Zarudny (1914) described the form in the Tashkent region during the breeding season, noted the narrow bars on the underparts and the rufous plumage phase in females, and found the birds were usually smaller than birds in Europe or Japan; Cuculus canorus bakeri Hartert, 1912; plumage above darker gray than in canorus and subtelephonus, underparts with black bars nearly as wide as white bars in some birds (range 1.2–2.6 mm), no rufous

plumage phase in adult females; SE Tibet, western China (W Sichuan), southern China to the Himalayan foothills in northeastern India, Assam, Nepal, Burma, Laos, northern Thailand and northern Vietnam; Cuculus canorus bangsi Oberholser, 1919; smaller, females often with rufous breast although there is no distinct barred rufous phase; Iberian Peninsula, Balearic Islands and the Mahgreb (Morocco, Algeria and Tunisia). Other subspecies have been described, but the difference between them and the forms recognized is slight and inconsistent between geographic regions. (1) Cuculus c. johanseni von Tschusi, 1903, described from Tomsk in western Siberia as smaller and paler than the European cuckoos (Portenko 1931, Johansen 1955, Malchevsky 1960), are intermediate in plumage between C. c. canorus and C. c. subtelephonus and intergrade across southern Siberia. von Tschusi named the form in 1903 after Professor Hermann Johansen. Hermann Johansen’s successor, Prof. Hans Johansen, collected these cuckoos, then left Siberia in 1937 and wrote his observations ( Johansen 1952, 1955); his collections are in ZMUC. (2) Cuculus c. telephonus Cabanis and Heine, 1863; Japan, smaller on average than European cuckoos, with larger pale spots in the tail and the underparts with narrower black bars; males with black bars (0.8–2.0 mm) narrower than that of males in Europe (reports of cuckoos in Japan with wider bars are misidentified Oriental Cuckoo C. optatus); (3) C. c. fallax Stresemann, 1930, in the southern coastal plain of China, birds are like C. canorus subtelephonus of western Sichuan in plumage with thin gray bars on the underparts, and the birds in the type series are within the size range of C. c. subtelephonus (wing, M (n ⫽ 2) 205, 210, and U (n ⫽ 1) 208 (ZMB)); the first two were reported as wing 204, 205 by Stresemann (1930a,b).Width of the bars on the underparts of the specimens are 1.2–2.0, 0.6–1.0 and 1.0–1.6 mm, much as in subtelephonus from central Asia to northern China. Other birds taken in the same locality in Yao Shan were the same size (wing, M (n ⫽ 5) 208–215) (Yen 1933). Although Mees (1979) reported that birds in coastal China are

512 Common Cuckoo Cuculus canorus similar in plumage to C. c. bakeri of eastern India, these birds in China are paler and more closely match the size, color and barring of C. c. subtelephonus; C. c. fallax is a synonym of C. c. subtelephonus. The geographic transition between dark C. canorus bakeri and pale C. c. subtelephonus is abrupt in eastern Tibet, where birds were sampled by Dolan and Schäfer (Stone 1933, Dolan 1939, Schäfer and Meyer de Schauensee 1939). Although they were identified in the collection report as the dark form C. c. bakeri, they in fact range from dark bakeri in eastern Tibet and western Sichuan (Songpan) where the black bars on the underparts are broader than 2.5 mm, to paler along the upper Yangtse basin at Jyekundo and camp 76, to pale subtelephonus at Wenchuan, western Sichuan, at the base of the Tibetan massif, where the black bars on the underparts are narrower than 1.6 mm (ANSP). Birds of western Yunnan near Tenggueh (AMNH) are variable, six cuckoos having bars less than 2 mm in width and the others with bars averaging about 2 mm. South of the Himalayas, cuckoos from Uttar Pradesh to Madhya Pradesh and Bombay are variable, with the extremes of dark bakeri to pale subtelephonus (Vaurie 1965, UMMZ). Birds taken recently in western Yunnan, western Sichuan and Qinghai, by CAS are recognized as C. c. bakeri. In Assam, C. c. bakeri are variable in bar width; all are dark on the upperparts.

Measurements and weights C. c. canorus, Europe: Wing, M (n ⫽ 52) 213–230 (221 ⫾ 4.3), F (n ⫽ 35), 204–216 (210 ⫾ 3.8); tail, M 204–216 (177 ⫾ 4.1), F 158–177 (167 ⫾ 5.46); bill, M 25.5–31.2 (27.7 ⫾ 15.1), F 25.5–31.2 (26.8 ⫾ 1.01) (Cramp 1985); Japan: Wing, M (n ⫽ 7) 213–226, F (n ⫽ 1) 220 (Yamashina 1941); M (n ⫽ 16) 203–227 (214.9 ⫾ 76.6), F (n ⫽ 3) 199–217 (205.7 ⫾ 9.9) (AMNH, FMNH, USNM); NE China (north of 30° N, most near Beijing): Wing, M (n ⫽ 17) 202–231 (219.1 ⫾ 8.9) (4 were ⬍ 210) (CAS); wing, M (n ⫽ 15) 206–228 (217.3 ⫾ 8.1) (4 were ⬍ 210), F (n ⫽ 9) 205–224 (213.0 ⫾ 5.6) (AMNH,ANSP, CM, FMNH, MCZ, UMMZ); C. c. subtelephonus, Turkmenistan (Tashkent region): Wing, M (n ⫽ 30) 201.3–249 (221.9 ⫾

15.2), F, gray phase (n ⫽ 18) 180–223 (200.5 ⫾ 13.1), F, rufous phase (n ⫽ 8) 180–217 (200.5 ⫾ 15.0) (Zarudny 1914); Transcaspia,Turkmenistan, Afghanistan and Iran: Wing, M (n ⫽ 30) 210–240 (223.5) (Paludan 1959, Vaurie 1965); Iran and Turkmenistan: Wing, M (n ⫽ 13) 200–241 (220.6 ⫾ 11.8), F (n ⫽ 6) 205– 224 (213.3 ⫾ 7.4), juvenile F (n ⫽ 1) 191 (FMNH); China, Xinjiang:Wing, M (n ⫽ 4) 218–230 (223.5), F (n ⫽ 2) 207–219 (213) (CAS); SE China (Hebei and Guilin regions): Wing, M (n ⫽ 15) 201–210 (205.5 ⫾ 3.4) (CAS); C. c. bakeri, Assam: Wing, M (n ⫽ 16) 198–232 (214.3 ⫾ 9.9), F (n ⫽ 9) 195–215 (205.1 ⫾ 6.8) (UMMZ, FMNH (F)); C. c. bangsi: Wing, M (n ⫽ 7) 199–215 (206.7 ⫾ 6.2), F (n ⫽ 5) 194–207 (200.6 ⫾ 5.4) (AMNH). Weight, Britain, May: M (n ⫽ 84) (117 ⫾ 17), F (n ⫽ 52) (106 ⫾ 14) (Seel 1977b); Camargue, France, May: M (n ⫽ 12) (101.0 ⫾ 8), F (n ⫽ 9) (96.0 ⫾ 9) (Glutz von Blotzheim and Bauer 1980); Japan: U (n ⫽ ?) 95–145 (119) (Enomoto 1941); Hopeh, NE China: M (n ⫽ 35) 71–127 (96.0), F (n ⫽ 14) 70–138 (97.0) (Shaw, in Cramp 1985); Iran: F (n ⫽ 1) 80.7 (Paludan 1938).Weights increase during the breeding season; before migration in late summer individual cuckoos increased from (1) 87 to 135 in 15 days, (2) 99 to 160 in 16 days, and (3) 98 to 149 in 17 days (Glutz von Blotzheim and Bauer 1980); by the time of migration, 150–197 (Löhrl 1979). Wing formula, P8 ⬎ 7 ⬎ 9 ⬎ 6 ⬎ 5 ⬎ 4 ⬎ 10 ⬎ 3 ⬎ 2 ⬎ 1.

Field characters Overall length 32–33 cm. Large gray cuckoo with a black bill, the lower breast and belly with black bars, under tail coverts white (buff in Oriental Cuckoo C. optatus, and nearly buff in some female C. canorus), either unbarred or with black bars, bend of wing barred black and white, the under wing with a broad white stripe across the secondaries and narrow across the primaries, not extending beyond P3–4. In Central and East Asia, C. c. subtelephonus and some C. c. canorus have black bars narrower than in C. optatus, but in western Russia C. c. canorus and C. optatus are equally barred below. Female in rufous plumage has rump with rufous bars broader

Common Cuckoo Cuculus canorus 513 than the black bars (black bars are broader in C. optatus). Juvenile has smaller white spots on the tail than juvenile C. optatus. Plumage is less broadly barred black below than in C. micropterus. In breeding season, distinguished by the two-note song, the first note louder and higher in pitch than the second. In Africa, juveniles differ from juvenile C. gularis in the presence of brown or tan bars in the plumage; in Asia, juveniles differ from the darker juvenile C. optatus in the same trait and are barred (not black) on the breast.

Voice Male song loud burry whistle “cuck-oo”, first note louder and higher, about 1 song/sec. Song length 0.5 sec, first note 0.12 sec, rises from 0.75 to 0.9 kHz (peak energy) then falls to 0.75 kHz, second note 0.18 sec, flat at 0.7 kHz. Songs are nearly constant across the Palearctic from Britain and Scandinavia to Japan and in India and Burma (Baker, in Bent 1940, Stresemann and Heinrich 1940, Rand and Fleming 1957, Ali and Ripley 1969, Cramp 1985, Payne 1986, Lekagul and Round 1991, Martens and Eck 1995, Hwang 1997, Chappuis 2000). Sometimes males give three-note songs (Thiede 1994). Excited male gives a harsh “gowk gowk gowk!”.The female gives a bubbling “kwik-kwik-kwik!”.

Range and status Palearctic, in Britain and Europe east to Siberia, China, Kurile Islands, Korea and Japan, south to Mediterranean Africa and the Middle East, southern Asia, India and southeast Asia (Yamashina 1931, Cramp 1985, Fiebig 1993, Ernst 1996, Shirihai 1996, Isenmann and Moali 2000,Yoon 2000). In the Indian subcontinent, they breed mainly in Baluchistan and northern Pakistan and in the Himalayas and hills of NE India (Grimmett et al. 1999). Seasonal migrant in the northern Palearctic, arriving in SW Britain over the sea from the south. In southern Yunnan, cuckoos of the breeding form C. c. bakeri occur from April through July or later, while the less broadly barred northern C. c. canorus are passage migrants in May (Bangs and Phillips 1914, MCZ, USNM). In hill and mountain country from Sikkim, Bhutan, Assam and Chin Hills to

SE Tibet and the Shan States C. c. bakeri are seasonal from March to August (Stevens 1925, Ludlow and Kinnear 1937, 1944, Smythies 1986), rare in northern Thailand (Lekagul and Round 1991), northern Laos and Tonkin (Bangs and Van Tyne 1931, Robson 2000a). Records in Taiwan apply to another cuckoo species, Himalayan Cuckoo C. saturatus (Mees 1979). C. c. bakeri are resident in tropical lowland areas of southern and southeast Asia, with few winter records and their wintering grounds are not well known, with records from the Maldives, Sri Lanka and the Andaman Is (Desfayes 1974, Hoffman 1989a). In Indonesia, an immature bird said to be C. canorus taken on islet Hoorn off West Java in January (Mees 1979, MZB 604), has the white carpals, large white underwing bar, and buff under tail coverts of C. optatus. European birds migrate flying nonstop more than 3000 km over the Mediterranean and the Sahara and winter in sub-Saharan Africa.Adults leave Europe in August and September, juveniles about a month later. A Dutch-ringed juvenile was taken in Togo in West Africa in October, and a British-ringed cuckoo was taken by bow and arrow in Cameroon in January. The distribution of ringing recoveries and museum specimens indicates that cuckoos in Britain and northern Europe migrate in a broad front from the Mediterranean, a few land on the northern edge of Africa, and most pass over the Sahara to winter in sub-saharan Africa (Seel 1977a, 1984a). Birds move along the east and west coasts of Africa and the Nile (Bannerman 1933, Moreau 1972, Gatter 1987, 1997, Zimmerman et al. 1996). Migrant cuckoos appear in Senegal from late July through October (Morel and Morel 1990), move southeast along the African coast in December and northwest in February and March (Gatter 1987, 1997). Nearly all records in West Africa are in autumn, and the few specimens reflect the lack of observation and collection in that region. G. D. Field and I observed a juvenile in Freetown, Sierra Leone, on 25 Dec 1973; its wing coverts had the brown and gray bars of C. canorus and not the gray plumage of C. gularis (Payne 1977a). Cuckoos winter in much of East and southern Africa (Moreau 1972, Lewis and Pomeroy 1989, Cheke and Walsh 1996, Zimmerman et al. 1996) and a few birds occur there in summer (Britton 1982). In southern Africa

514 Common Cuckoo Cuculus canorus

X

?

C. c. canorus arrive in October and November and leave by April, and some are fat on their southern wintering grounds in December (Payne 1968, Vernon et al. 1997). C. c. bangsi occur in passage and winter in West Africa (Mt Nimba in Liberia, Ivory Coast, Ghana, Cameroon) (Moreau 1972, Colston and Curry-Lindahl 1986, Louette and Herroelen 1993, BMNH, LACM). Small cuckoos (females, wing 192, 197, ZMB) taken in November, February and March from Ituri and Kivu south along the west shore of Lake Tanganyika were identified by Stresemann (1928) as C. c. bangsi, as was a female (wing “178”) in November at Lupiala, Shaba (IRSNB:Verheyen 1953). However, there is no evidence of movement from West Africa across Zaire to SE Africa (Chapin 1939, Seel 1984a, Louette and Herroelens 1993) and these birds are the same size as breeding C. c. subtelephonus in Asia. Another small female (BWYO 17517) in rufous plumage with wing 197 was taken southeast of Lake Tanganyika at

Rukwa,Tanzania, on 15 Dec 1953; a male (FMNH 81681) with wing 200 was taken at Katungulu, Mwanza, south of Lake Victoria in northern Tanzania, on 23 Feb 1927, and a female (UMMZ 216723) in rufous plumage with wing 198 was taken at Alexandria Forest, Eastern Cape, South Africa, on 21 Oct 1966. These birds might be small northern European C. c. canorus (suggested to be Zaire birds by Louette and Herroelens 1993); more likely the wintering cuckoos of the rift region are Asian C. c. subtelephonus (which they match in size and in the rufous female plumage, and which winter in East Africa). Asian C. c. canorus and C. c. bakeri winter in India and southeast Asia. C. c. subtelephonus winter in India and in East Africa, arriving in Somalia from 3 Sept to 10 Nov and appearing later in the season in Kenya and Uganda and eastern Zaire to Tanzania and N Mozambique, rarely in southern Africa to S Mozambique, E Zimbabwe and KwaZulu-Natal (Cramp 1985, Clancey 1961, 1980, 1996, Irwin

Common Cuckoo Cuculus canorus 515 1981, Ash and Miskell 1998; BMNH, MCZ, ZMUC, ZSM). Cuckoos are silent on their winter grounds. They are quiet moving north along the coast in East Africa from March to April (Stevenson and Fanshawe 2002), but they sing during spring migration in Somalia from late April to late May (Ash and Miskell 1998) and in the rift valley while feeding on the way ( Jackson 1938); migrants do not sing in West Africa. Migrants occur in passage in the Middle East, and some may breed there. Migrant C. c. subtelephonus appear on islands in the western Indian Ocean (Seychelles, Aldabra, îles Crozet, Marion I) (Seel 1984a, Stahl et al. 1984, Irwin 1989, Betts 2002) and in Oman and the United Arab Emirates (Gallagher and Woodcock 1980, Richardson 1990). In Japan, males return to their breeding grounds in mid-May, females a week later (Yoshino 1999). C. c. canorus are accidental in Iceland and other islands in the North Atlantic (Faeroes, Azores, Madeira, Canary Islands and Cape Verde Islands), and rare in Alaska and eastern North America (Cramp 1985, AOU 1998). In cuckoos ringed in the breeding season and recovered a later year, 25% of 12 adults were found on the same site, and 50% were within 20 km of the ringing site (Lindholm 1999). Males marked as nestlings returned to their natal area in the next season. One bird with a few retained juvenile feathers was territorial, sang, and interacted sexually with a female cuckoo that gave a bubbling call, and a female with a few old juvenile feathers had an egg in the oviduct (Seel et al. 1981). Survival from nestling to fledging is 22% (Wyllie 1981), survival of fledged juveniles to end of first year is c. 28%, and annual survival of adults is c. 52% (Seel, in Brooke and Davies 1987). The breeding range of a laying female using host Reed Warblers Acrocephalus scirpaceus was c. 30 ha; a female using Meadow Pipits Anthus pratensis was active within c. 300 ha (Wyllie 1981, Hagenmeijer and Blair 1997). Population density varies with habitat; some males call 5 km apart (Dröscher 1990), estimated average density 1–2 males/km2 in northern Europe (Cramp 1985,Viksnes 1989) and as many as 20–30 birds/km2 in Spain (Dias et al. 1996). Population estimates, 25,000–35,000 pairs in Latvia, fewer than

10,000–35,000 pairs in Britain and Ireland, and 100,000–1,000,000 pairs in France; numbers decreased in northern and western Europe during the twentieth century (Cramp 1985, Brooke and Davies 1987, Strazds 1996). Population in Europe estimated at 1,560,000 pairs, in Russia 700,000 pairs, in Turkey 160,000 pairs (Hagenmeijer and Blair 1997). Breeding density near Mussoorie, India, as many as 5 singing males along five miles (c. 1 bird per km2) (Fleming 1967).

Habitat and general habits Forests, both coniferous and deciduous in Kashmir, rarely in closed forest, more often in second growth, open wooded areas, wooded steppe, scrub, agricultural lands with scattered trees, and treeless meadows; lowlands and hill country to 2000 m (Stevens 1925, Cramp 1985, Price and Jamdar 1990, Ernst 1992a). In Iberian Peninsula the highest densities are in oak woods (Quercus petraea, Q. pyrenaica) (Dias et al. 1996). In Nepal the cuckoo is one of the most wide-ranging birds, from woodlands to brushy vegetation above timberline and subalpine meadows, from 500 m to 3000 m (and to 4250 m) (Martens and Eck 1995), and in India they occur as high as 5250 m (Grimmett et al. 1999). Below sea level at the base of mountains in Tulufan Depression, Xinjiang, western China, this is a common breeding bird (CAS, Lei Fu-min). Birds wintering in Africa avoid forest habitats.These cuckoos are mostly solitary, and they are inconspicuous when not singing; occasionally they occur in small groups in winter (Irwin 1988).

Food Insects, mainly caterpillars (notably armyworms Spodoptera exempta), less often dragonflies, damselflies, mayflies, crickets, cicadas, paper wasps and bees, bugs, beetles in cold weather; spiders, snails; rarely fruit. Females take eggs and nestlings of small birds (Gurney 1897, Löhrl 1979, Cramp 1985, Roberts 1991, Hagenmeijer and Blair 1997).

Displays and breeding behavior Males are territorial, singing consistently in the breeding season. Radio-tracked males in Europe

516 Common Cuckoo Cuculus canorus and Japan have territory areas of several square kilometers. Males feed to 4 km from the singing site, and females feed up to 2–3 km from their laying sites (Dröscher 1990, Nakamura and Miyazawa 1990, 1997). In courtship behavior, either the male or female may initiate courtship with calls, the male with advertising call and the female with bubbling call. The male perches near the female, bobs head, bows, open and droops wings, and raises and fans tail, rotates his body around the horizontal axis and sways from side to side (Wyllie 1981). Courtship feeding is seldom seen. Male also circles in a display flight and calls to female. In mating, male follows female quietly, female opens her wings slightly and moves her tail aside and calls, the male mounts, moving his tail from side to side (Molnár 1944,Wyllie 1981). Copulation occurs in the morning, and the fertilized egg is laid the next afternoon (Chance 1922, Gärtner 1981a,b). A female often mates with only one male in a season ( Jones et al. 1997).

Breeding In Britain and elsewhere in NW Europe they breed in May and June, in Algeria in April and May (Isenmann and Moali 2000), in India and Burma from March to August (Baker, in Bent 1940,Ali and Ripley 1969, Smythies 1986, Ali 1996). In Japan, these cuckoos breed from late May to mid-July or later (Yoshino 1999). Brood-parasitic. Hosts are insectivorous songbirds. In Europe and India more than 100 species have been noted with a cuckoo egg in the nest, including flycatchers, chats, warblers, wagtails and buntings (Baldamus 1853, 1892, Rey 1892, Capek 1896, Stresemann 1934, Makatsch 1937, Baker 1934, 1942, in Bent 1940, Dybbro 1977, Wyllie 1981, Brooke and Davies 1987, Mal’chevskii 1987, Marchetti 1992). In Britain the main hosts are Dunnock Prunella modularis, Reed Warbler Acrocephalus scirpaceus, Meadow Pipit Anthus pratensis, European Robin Erithacus rubecula, and White (Pied) Wagtail Motacilla alba (Brooke and Davies 1987). In India they parasitize mainly chats, also pipits and buntings (Baker 1942). In Algeria they use Sardinian Warbler Sylvia melanocephala, Tristram’s Warbler S. deserticola and Moussier’s Redstart Phoenicurus moussieri (Chavigny and Le Dû 1938).

In Japan, cuckoos are known to parasitize 20 species of host (Higuchi 1998, Nakamura et al. 1998), and in Korea their hosts include Daurian Redstart Phoenicurus auroreus and Vinous-throated Parrotbill Paradoxornis webbianus (Yoon 2000, pers. comm.). Each female cuckoo uses one species of host (Chance 1922, Marchetti et al. 1998, Gibbs et al. 2000). She removes and eats a host egg from the nest when she lays her own, or she removes the egg earlier in the day. Cuckoos removed eggs from c. 30% of the nests of Marsh Warblers Acrocephalus palustris) that they did not parasitize; the warbler abandoned 64% of these and often built a new nest which the cuckoo then parasitized (Capek 1896, Selous 1912–13, Chance 1922, 1940, Jourdain 1925, Gehringer 1979, Gärtner 1981a,b, Wyllie 1981, Cramp 1985, Yoshino 1988, Matsuda and Uchida 1990, Alvarez 1994a). Occasionally, seedeating finches (Chaffinch Fringilla coelebs and carduelines including Greenfinch Carduelis chloris) bring insects to their young cuckoos even though the finches rear their own young on seeds (Glue and Morgan 1972, Seel and Davis 1981). Cuckoo eggs are polymorphic in color (blue, pink, whitish) and pattern (spotted or unmarked) in Europe, where 15 egg morphs are recognized, and about 77% of eggs found in a nest closely match the host eggs in color and pattern (Øien et al. 1995), although in some regions including Hungary the match is not very close (Moskát and Honza 2002). In Algeria the cuckoo eggs mimic the eggs of Moussier’s Redstart (Chavigny and Le Dû 1938). Cuckoos mimic eggs of several hosts in eastern India (Baker, in Bent 1940, Ali and Ripley 1969). In the Shan States, Burma, they lay spotted eggs in nests of Pied Bush Chat Saxicola caprata and blue eggs in nests of Dark-gray Bush Chat S. ferrea (Osmaston 1916, Livesey 1933, 1936, 1938a,b, Baker 1942). In Japan the eggs are spotted, marked with streaks, fine spots, irregular markings and sometimes lines, even though most hosts have no lines; cuckoo eggs laid in nests of hosts that rear their young are not similar to eggs of their hosts, and cuckoos lack the distinct egg morphs of cuckoos in Europe (Higuchi 1998, Nakamura et al. 1998). Eggs 23 ⫽ 1 7 mm in Europe (Cramp 1985), 23.4 ⫽ 17.7 mm in Japan (Higuchi 1998).The egg is

Common Cuckoo Cuculus canorus 517 usually laid when a host nest has 1 or 2 eggs (Chance 1922, 1940,Wyllie 1981).The female lays by perching on the nest, or at inaccessible hole nests by lifting herself against the nest with outspread wings and tail and ejecting her egg into the hole. She lays on alternate days (Chance 1922, 1940, Witherby 1938, Seel 1973). She takes only 8-20 sec to arrive at the nest, lay an egg and leave the nest. Incubation period is 11.5-12.5 days. A female lays as few as 8 or as many as 25 eggs in a season, depending on the availability of host nests (Chance 1922, 1940, Blaise 1965, Cramp 1985, Bayliss 1989a,b). The nestling cuckoo is 2.6 g at hatching, 4-5 g at day 2, 9-12 g at day 4, and 41-46 g at day 9, and it fledges at 80-100 g (Bussmann 1947, Olsson 1948).The nestlilng evicts the eggs of the host from hatch day to 2 days after hatching; in Japan it takes a few days to push out the eggs and nestlings of a large host Azure-winged Magpie Cyanopica cyanea. The nestling period is 17-18 days in Britain and 19-20 days in Sweden, Germany and Japan. Beg-

ging calls are the same in nestling cuckoos in the nests of different species of hosts; there appear to be no host races in begging behavior (Butchert et al. 2003). After fledging, the young are fed by their foster parents for another 2-3 weeks (Witherby et al. 1938, Bussmann 1947, Werth 1947, Olsson 1948, Wyllie 1981, Cramp 1985, Alvarez 1994b, Yoshino 1999). A nestling cuckoo’s demand for food and parental care is about the same as that of a brood of the host’s own young (Brooke and Davies 1989, Davies et al. 1998). Breeding success of cuckoos was 56% in a population in Norway, where Meadow Pipit was the main host. In this region, 6.4% of pipit nests were parasitized, to 25% in some low-lying sites (Moksnes and Røskaft 1987). In one area in Hungary, 665 of 2073 (32%) nests of European Robin were parasitized (Varga 1994). In Britain, only 1.4% of the nests of the most commonly parasitized host species have a cuckoo egg; locally as many as 56% of Reed Warbler nests have a cuckoo egg (Brooke and Davies 1987).

This page intentionally left blank

Glossary

adaptation a trait that is an evolutionary solution to an ecological challenge, sometimes referred to as a “key innovation” in a clade allopatric species or populations that occur in different geographic areas amplitude the loud part of a song or call apomorphy a trait that arose in a descendant lineage, a derived trait apterium a bare patch between the feather tracts audiospectrogram a visual graph of a call or song, with time on the horizontal axis and frequency (or pitch) on the vertical axis austral a geographic term referring to the southern hemisphere Australasia the geographical region extending from southern Asia to Australia barb side branches from the shaft of a feather, the branches linked to form the vane of the feather barbule side branches from the barb of a feather basal a taxon that branches off an evolutionary lineage near the base basipterygoid process a lateral process extending from the ventral surface of the skull to the pterygoid biological species the set of individuals that interbreed with each other (allowing for sex), sometimes recognized by a common song bloom a dull coat of particles that overlays a shiny surface, as in ripe grapes, optical lenses and the chalky layer over a shiny layer on bird eggs bootstrap a “bootstrap” analysis in systematics resamples the set of characters for all taxa; it

uses the same number of characters as in the original data set. Each replication involves a random sample of the characters, sampling with replacement, such that any one character can be sampled more than once or not at all in a given bootstrap replicate.The higher the bootstrap value, the higher the proportion of these replicate samples found in the same indicated clade. Because the replicates are drawn from the same data and use a high proportion of the same characters in each test, they are not independent samples. In spite of this caveat, statistical probability has often been inferred from these bootstrap values. boreal a geographic term referring to the northern hemisphere broad band a sound that covers a wide range of frequencies at any one time brood parasite a bird that lays its eggs in the nest of another bird, and uses the parental care of the other species (the host) to rear its young; an obligate brood parasite never cares for its own young brood patch a bare area on the ventral surface of a breeding bird that becomes thick and vascular in the breeding season and transfers heat from the adult to the incubated egg carotenoids red and yellow pigments that are produced by plants; incorporated in bird plumage when taken in the diet and stored in growing feathers catapophyses paired ventrolateral projections on the anterior end of the vertebral centrum, that extend lateral to the prezygapophyses and

520 Glossary involve the carotid; the Proc. ventrolaterales of Baumel (1993) choana a slit or opening in the roof of the mouth choana, pars rostralis a narrow rostral slit in the anterior palate choana, pars caudalis a triangular opening caudal to the maxillopalatine clade a group of animals that have a common ancestor; a branch on the evolutionary tree of life (after the Greek “klados”, “branch”) cladistic a logic of estimation of the relationships among taxa that determines the smallest number of evolutionary changes that were necessary to account for the distribution of characters among these taxa from a common ancestor cladogram a tree diagram of branching sequences of evolutionary relationships in which the relative genetic distances between taxa are not indicated (cf phylogram) clutch size the number of eggs laid in a nest; e.g. 2(3) indicates usually 2, sometimes 3 eggs; by extension, the number of eggs laid in a time series by a female brood parasite

description a set of words or illustrations that characterizes a taxon desmognathous bony palate with the maxillopalatines in contact in the midline, and the pterygoids and palatines in articulation with the basisphenoid rostrum (vs schizognathous) diagnosis a summary of characters that differentiate a taxon from related or similar taxa (see “description”) dimorphism difference in form, used mainly in the sexual context, i.e. between male and female; sometimes used for difference in forms of immature and adult dispersal the movement of individuals from one area to another, either breeding dispersal of adults (movement between successive breeding sites), or natal dispersal (movement from the natal site to the breeding site) DNA the molecular code of genetics, with the elements being the four nucleotide bases (A, C, G,T)

coevolution the evolution of adaptations between interacting species, especially a reciprocal evolutionary interaction such as the arms-race of cuckoo egg mimicry and host egg recognition

DNA-DNA hybridization a technique of molecular genetics that compares genetic similarity between strands of DNA by radiolabeling certain tracer strands, combining (“hybridizing”) strands into duplex strands by cooling, and determining the rate of thermal disassociation of the duplex strands with a series of increases in temperature (Sibley and Ahlquist 1990); the technique estimates the degree of relationship between species

condyle a rounded process at the end of a bone, articulating with another bone

egg-dumping laying an egg by one female in the nest of another female

consistency index, CI the percentage of evolutionary changes in individual characters that are consistent with a cladistic tree of taxa

endangered a conservation term referring to a species living in danger of extinction, either because it occurs in small numbers, or because it is confined to a small geographic range and is under threat of further loss of population

codon a triplet of nucleotides that represents an amino acid or a termination signal

cooperative breeding the occurrence of mature birds in rearing young that are not their own offspring, when living in a social group courtship feeding feeding of female with an insect or ripe by a male, possibly resulting in copulation

endemic geographical distribution of a taxon restricted to one area

culmen the dorsal ridge of the upper mandible

eutaxic wing with equal numbers of secondaries and greater secondary coverts

derivative a character or trait that was not present in the ancestor of a clade, but evolved later in the lineage

evolution change through time, especially genetic and morphological change over many generations, within current populations as well as

Glossary 521 among species derived from a common ancestor; used in studies of phylogeny and of adaptation extinct a species or lineage that once lived but now has no living members foster to provide parental care to a young bird (such as a brood-parasitic cuckoo) that is not the birds, own offspring gape the base or corner of the mouth, especially of a young bird where this area is large and brightly colored gene a unit of heredity; a sequence of nucleotides in DNA, the term often restricted to a transcribed sequence of nucleotides genetic distance the proportion of nucleotide differences between corresponding genes in two taxa genotype the genetic constitution of an individual gens, pl. gentes kind, especially in egg morphs of polymorphic brood-parasitic cuckoos, where the genetic basis indicates a race, at least in theory (gens theory). Because only the female shows the trait, these races may be specific to the sex chromosome carried only by the female, and need not involve their mates.Although “gens” applied to cuckoos refers to a race (of females) that lays a similar kind of egg and lays in the nest of the same host species, the term dates back to Greek and Latin “gens” relating to a clan, tribe, or family. Plato used the term γε´νoσ interchangeably with ειδoσ to indicate a group or its parts;Aristotle used the term with no consistent distiction both as a collective set and as its members or elements (Balme, D. M. 1962. ΓΕΝOΣ and EI∆OΣ in Aristotle’s biology. Classical Quarterly, new series, 12, 81–98) — much as “species” is used at the present time in biology to indicate a unit (phylogenetic species) or its members (biological species). It is used in an inclusive social sense by English social classes (“Gentle folk”) and the Pope (“Decree Ad Gentes”, to the people of God, 1965), and by the egalitarian French (with a meaning of people, or nations).The term has social implications that shed light on its use in our birds. Marxist theoreticians used the term to describe a social institution, in a historical sense of social evolution

(Engels 1972, The origins of the family, private property and the state, translation). In this view, long ago, people lived in communal families, or maternal gentes.These groups were clans (this term later was taken over to refer to a patrilineal kinship lineage) and these maternal gentes were an earlier stage of the patriarchal gens of civilized people.The “gens” principle had “the same importance for anthropology that Darwin’s theory of evolution has for biology, and Marx’s theory of surplus value has for political economy”. Engels applied the term from Roman Law to promote the view that a matrilineal “gens” was the ideal of social organization, prescribed ownership and prescribed sexual access. In a genealogical sense, a gens referred to all individuals who traced their ancestry to the same mother. In this structure of family, only the female was important. In this view, matrilineal “gens” has passed in transition in “civilized” societies to patrilineal lineage.The patrilineal “clan” or “gens” of the Romans was then a social convention that consolidated group ownership of production into the hands of a capitalist class, and this shift of family organization had the consequence of taking females out of the lines of perceived social power and production in labor. In cuckoos, Baldamus (1856, 1896) developed the idea that individual females lay an egg of a certain color and pattern, usually in the nest of a host species that lays a similar egg, and the implication of matrilineal races where a mother has a daughter with a similar egg. Newton (1896) did not use the “gens” term; Southern (1954) established the term in the cuckoo literature. genus, pl. genera, a clade of closely-related species, the scientific name to indicate phylogenetic relationships among species gonys the ventral ridge of the lower mandible ground-cuckoo one of the genera of large cuckoos Neomorphus or Carpococcyx ground cuckoo a cuckoo that lives on the ground; in the New World a cuckoo in the subfamily Neomorphinae group a set of individuals or units; a social set of interacting individuals

522 Glossary haemuli hooks on the feather barbule that connect it to the adjacent barbule and hold the vane in shape harmonic an integral multiple of a fundamental frequency in a call or song, evident as a trace on the audiospectrogram helper an individual that provides care to young of other individuals during the period of parental care of the young heuristic approach approximate methods used when the complexity of a problem prevents all possible solutions from being considered holorhinal nostrils without a deep cleft of the bony structure on the posterior edge holotype the unique specimen upon which the original description of a taxon was based homology a trait that appears in lineages as a result of common ancestry homoplasy a trait that appeared independently in different lineages and was not present in the common ancestor of the lineages host a species that rears the young of another species which is its brood parasite hybrid individual resulting from a mating between different species hypapophysis ventral ridge on the vertebral centrum; the Proc. ventralis of Boas (1929) and Baumel (1993) hypotarsus crest on the ventral surface of the proximal end of the tarsometatarsus ingroup the set of taxa of interest in a phylogenetic study juvenile a young bird in its first plumage after leaving the nest; includes young birds that feed independently keratin a protein structure which makes up the feathers and scales in birds kHz kilohertz, a unit of pitch or frequency in sound, where 1 kHz = a thousand cycles per second kin selection the form of natural selection where the behavior of an individual affects the success

of close relatives, especially those that are not direct descendants of the individual (that is, not its offspring); also known as indirect selection lineage a group of species or genes related by descent from a common ancestor lores the area between the eyes and the base of the upper half of the bill Malay Archipelago a geographic region from the Malay Peninsula through Indonesia, Borneo, Sulawesi and the Moluccas, excepting New Guinea; the term used mainly historically by A. R.Wallace, the region (in part) also called Sundaland Melanesia the Pacific island region east and south of New Guinea, including the Admiralty, Bismarck and Solomon archipelagos, Fiji, Vanuatu and New Caledonia. melanin a dark brown or black pigment that occurs in feathers, skin and other tissues and that is polymerized from tyrosine microsatellite a tandem repeated base sequence in a fragment of DNA, where the bases are repeated in variable numbers between individuals (for example,ATAT vs ATATATAT). Homologous microsatellites are identified by their position in relation to a conserved genetic marker using a flanking primer. Comparison of these sequence lengths between individuals for several such repeats allows matching of genetic similarity, parentage and kinship of these individuals. mimicry the resemblance of one species with another (the model), where the resemblance gives the mimic the advantage of being recognized as the model; and is not due to common descent miombo a seasonally deciduous forest type in central Africa, dominated by leguminous trees Brachystegia and Isoberlinia, the trees c. 20 meters high with leaves emerging before the rains mitochondrial DNA the genetic constitution of an extranuclear organelle that is transmitted from mother to offspring; mtDNA copies are present in large numbers and allow molecular genetic comparisons of individuals and species in estimating their evolutionary relationships

Glossary 523 monogamous a mating unit of one male and one female, where each sex has a single mate; a species where this breeding unit is the most common mating system monophyletic a genetic or evolutionary lineage that has a single common ancestor and includes all the branches that were derived from this common ancestor monotypic any higher taxon that comprises only one set or lower taxon, i.e., a genus of only one species; a species with no subspecies morph a form or a phase in a population morphological species a set of populations that is distinct in visual appearance from other such sets muscle formula letter codes indicating the presence or absence of certain thigh muscles museum a collection of specimens for research, a center for study and research, sometimes available on exhibition for nonspecialists

parapatric occurring in contiguous neighboring geographic areas, rather than separated or overlapping areas paraphyletic a set of lineages that do not include all the current lineages that were derived from their common ancestor, and is not comprehensively monophyletic parsimony a logic that involves the fewest assumptions or the smallest number of changes passage migrant a population that occurs regularly in an area between its breeding grounds and its wintering grounds passerine a member of the order Passeriformes, including the songbirds; the passerines include the host species of brood-parasitic cuckoos phenetic a character similarity that is compared among species without regard to the uniqueness of its evolution in a clade

nucleotide the units of molecular genetics, the four bases (adenine, cystine, guanine, tyrosine)

phylogenetic the relationship among species based on their branching order from a common ancestor; the logic of reconstructing evolutionary relationships based on the set of characters that differ from a common ancestor, or its proxy a related “outgroup”; the resulting estimate of phylogenetic relationships includes both the branching order and the number of characters that distinguish each clade (vs “cladistics”, which is concerned strictly with the branching order)

outgroup a species or set of species that are related to another set of species whose relationships are the object of study (the “ingroup”); a reference group that serves as a proxy for characters shared by the common ancestor of an ingroup

phylogenetic species a species defined in terms of its ancestry, a distinct evolutionary lineage (operationally defined as a morphological species that is discontinuous in characters from other such species)

overtone a non-integral or an integral harmonic of a fundamental frequency in a call or song, evident as a trace on the audiospectrogram

phylogeny the true evolutionary history of a group; also used in the sense of an estimate or a diagram representing the branching evolutionary history of a set of taxa

Neotropical relating to the New World tropics; the biogeographic region from Mexico south through South America node the point of branching in a tree notch a V-shaped mark of contrasting pigment color on the margin of a feather; the feather itself is not indented

palate the roof of the mouth, particularly the skeleton including the form of the palatine and maxillary bones Papuasia a subregion of Australasia comprising the Moluccas, New Guinea and northern Melanesia

phylogram a diagram of evolutionary relationships in which the relative genetic distances between taxa are indicated (cf cladogram) plesiomorphy a trait that was present in the common ancestors of a lineage

524 Glossary polyandry a mating unit of one female and more than one male (applies to a single mating unit group, and to a population or species where this is the common mating system)

scutes the scales on the foot (exposed tarsometatarsus and toes)

polygyny a mating unit of one male and more than one female (applies to a single mating unit group, and to a population or species where this is the common mating system)

secondary growth new forest growing where an earlier original forest was removed

polymorphic condition in which two or more forms are present in a population polyphyletic a set of lineages that do not share an exclusive and common ancestor polytypic more than one kind; in a species refers to distinct appearances in different geographic regions, especially where the forms do not intergrade and the populations are named as distinct subspecies primaries the large flight feathers of the outer wing primitive present in the common ancestors of a lineage, symplesiomorphy pterylography the arrangement of feather tracts on the skin of a bird race a genetic variation within a population, not necessarily based on geographic region (see subspecies)

secondaries the flight feathers of the wing that are attached to the ulna

sehnenhalter an accessory articular flange or process on trochlea IV of the tarsometatarsus, occurring in zygodactyl birds sequence the linear ordering of bases in DNA sex chromosome chromosomes that differ between males and females. Among birds, males are homogametic with pairing chromosomes (ZZ) and females are heterogametic (ZW) sexual dimorphism the difference between males and females in plumage, size, or other characteristics sexual selection the process of evolutionary change that is due to the individual advantage gained in obtaining a mate, independent of any advantage in survival septum a dividing shelf shaft the axis of a feather or bone

rattan a thorny climbing palm

sister group the clade that shares a most recent common ancestor with another clade; sister species are each other’s closest relatives

rectrices the large flight feathers of the tail (singular, rectrix)

songbird the oscines, the major lineage of passerines

remiges the large flight feathers of the wing (singular, remix), the primaries and secondaries

specialized a population, species or set of species that is distinct and has traits suited to a particular life style

rictal bristles stiff or narrow feathers at the corner of the mouth, without fused barbs, and often directed forward along the bill rooting in a cladistic study, the position in a tree at which the ingroup is associated with the outgroup; the ancestral rooting for an ingroup sometimes varies with the reference taxa that are chosen as the outgroup scapulars the dorsal feathers at the base of the wing schizognathous bony palate with the maxillopalatines not meeting in mid-line with each other or with the vomer (vs desmognathous)

speciation the process through which one species splits into two descendant populations that are genetically, morphologically and behaviorally distinct and do not interbreed with each other species a distinct kind, a population that interbreeds within itself and does not interbreed with other kinds; often recognized by morphological distinctiveness species pair two species that are each other’s closest relatives; sister species subcanopy the canopy of forest trees that mature at midstory level

Glossary 525 subspecies recognizable and taxonomically named distinct populations of a single species that live in different geographic areas

tracheobronchial anatomically involving both the trachea and the bronchi

sulcus groove

transition in a phylogenetic analysis, the shift of character state

superspecies a monophyletic set of closelyrelated allopatric species

tree a rooted network with no reticulation; the branching sequence in a phylogeny

sympatric occurring in the same geographical area, contrasting with allopatric and parapatric

trochlea a bony process with a groove that guides movement of a tendon or another bone

symplesiomorphy primitive character; a trait that is present in descendants as a result of the presence in an ancestor (cf. synapomorphy)

tropical the geographic region between the tropic of Cancer and the tropic of Capricorn

synapomorphy derived shared character; a trait that is present in a set of descendants as a result of its origin in a clade from a common ancestor that lacked the trait (cf. symplesiomorphy) synsacrum the fused bones of the pelvis syrinx the enlargement of the trachea and bronchi, which is a sound-producing organ. “Syrinx” was a mythical water nymph who metamorphosed into a reed to escape the amorous attentions of Pan. Pan plucked the reed to make his pipe; her transformation was in vain. systematics the logical study of relationships, especially the evolutionary pattern of origin and descent of species tarsus the foot of a bird, especially when measured as the tarsometatarsus (proximal end of the tarsometatarsus to the distal end of the last undivided scute over the toes)

tuberosity an elevated surface, generally smooth and round type series the series of specimens upon which the original description of a form was based underparts the plumage on the ventral surface of the bird, from chin through belly and under tail coverts vagrant a bird of irregular or accidental occurrence in an area which is far from its usual breeding or wintering ground and is not on the usual migration route of the species vicariance the process of origin of a stand-in or proxy for another, especially a geographic counterpart that results from a geological separation of a formerly continuous set of populations vomer a small bone in the midline of the palate W-chromosome the chromosome found in female birds and not in male birds

taxon (plural, taxa) a set that has been named in classification

Wallacea a geographic region including Sulawesi, the Moluccas and Lesser Sunda Islands

taxonomy the classification of different kinds of animals

Wallace’s line a biogeographic discontinuity between the continental shelf of Asia on one side, and Sulawesi, the Moluccas, the Lesser Sundas the regions of New Guinea and Australia on the other side

territory the area defended by an individual bird, a breeding pair or a social group threatened a conservation term of assessment of populations that are small or declining, but are not as likely to undergo rapid extinction as those endangered tibiotarsus the bone in the leg that is formed by the tibia and the proximal tarsal elements that are fused to it; corresponds to the tibia of mammals

weighting the relative importance assigned to different characters in a phylogenetic analysis wing formula the relative length of the primaries in the folded wing zygodactyl arrangement of toes with numbers 1 (inner) and 4 (outer) directed backward, 2 and 3 directed forward

Bibliography

Aagaard, C. J. (1930). The common birds of Bangkok. Chr. Backhausen, Copenhagen. Abdulali, H. (1931). Eleven Koel eggs in a crow’s nest. Journal of the Bombay Natural History Society, 35, 458. Abdulali, H. (1967). The birds of the Nicobar Islands, with notes on some Andaman birds. Journal of the Bombay Natural History Society, 64, 139–90. Abdulali, H. (1971). A catalogue of the birds in the collection of the Bombay Natural History Society—10, Cuculidae. Journal of the Bombay Natural History Society, 68, 756–72. Abdulali, H. (1977). The southern limits of the Himalayan Cuckoo Cuculus saturatus saturatus (Blyth). Journal of the Bombay Natural History Society, 74, 172–3. Aboitiz, F. (1996). Does bigger mean better? Evolutionary determinants of brain size and structure. Brain, Behavior and Evolution, 47, 225–45. Abouheif, E. and Fairbairn, D. J. (1997). A comparative analysis of allometry for sexual size dimorphism: assessing Rensch’s rule. American Naturalist, 149, 540–62. Abrahamovich, A. H. and Cicchino, A. C. (1985). Estudios bioecologicos, sistematicos y filogeneticos de los malofagos parasitos de Guira guira Gmelin (Aves, Cuculidae): Vernoniella bergi (Kellog, 1906) (Philopteridae) y Osborniella guiraensis (Kellog, 1906) (Meooponidae). I. Identificacion de los huevos. Historia Natural (Mendoza, Corrientes, Argentina), 5, 209–16. Achard, F., Eva, H. D., Stibig, H.-J., Mayaux, P., Gallego, J., Richards, T., and Malingreau, J.-P. (2002). Determination of deforestation rates of the world’s humid tropical forests. Science, 297, 999–1002.

Ackerman, M. (1841). Note sur le coua, famac-acora des Malagaches, hache-escargot (traduction littérale) ou casseur d’escargots. Revue Zoologique, par la Société Cuvierienne, 4, 209–10. Alexander, B. (1902). On the birds of the Gold Coast Colony and its hinterland. Ibis, Ser. 8, vol. 2, 278–333, 355–77. Agro, D. (1994). Grasshoppers as food source for Black-billed Cuckoo. Ontario Birds, 12, 28–9. Åhlund, M. and Andersson, M. (2001). Brood parasitism: female ducks can double their reproduction Nature, 414, 600–1. Ali, S. A. (1927). The Moghul emperors of India as naturalists and sportsmen, part 2. Journal of the Bombay Natural History Society, 32, 34–63. Ali, S. (1931). The origin of mimicry in cuckoo eggs. Journal of the Bombay Natural History Society, 34, 1067–70. Ali, S. (1946). The book of Indian birds. (4th edn.) Bombay Natural History Society, Bombay. Ali, S. (1949). Indian hill birds. Oxford University Press, Bombay. Ali, S. (1953). The birds of Travancore and Cochin. Oxford University Press, Bombay. Ali, S. (1962). The birds of Sikkim. Oxford University Press, Bombay. Ali, S. (1996). The book of Indian birds. (12th edn.) Bombay Natural History Society, Bombay. Ali, S. and Ripley, S. D. (1948). The birds of the Mishmi Hills. Journal of the Bombay Natural History Society, 48, 1–37. Ali, S. and Ripley, S. D. (1969). Handbook of the birds of India and Pakistan, vol. 3. Oxford University Press, Bombay.

Bibliography 527 Ali, S. and Whistler, H. (1937). The ornithology of Travancore and Cochin, part 6. Journal of the Bombay Natural History Society, 39, 1–35.

AOU (American Ornithologists’ Union). (1998). Check-list of North American birds. (7th edn.) American Ornithologists’ Union, Washington, D.C.

Allan, D. G. (2002). First record of Striped Cuckoo Clamator levaillantii parasitizing Pied Babbler Turdoides bicolor. Bird Numbers, 11(2), 30.

Anciaux, M.-R. (2000). Approche de la phénologie de la migration des migrateurs intra-africains de l’intérieur des terres du sud-Bénin (plateau d’Allada et sud de la dépression de la Lama). 1. Les non-Passériformes et les non-Coraciiformes. Alauda, 68, 311–320.

Allen, J. (2003). Observations of a male Common Koel at Magill and Wattle Park. South Australian Ornithologist, 34, 88–9. Alley, T. H. (1978). Observations on the diet of a fledgling Rufous-tailed Bronze-cuckoo. Corella, 2, 75. Allport, G. A. and Fanshawe, J. R. (1994). Is the Thick-billed Cuckoo Pachycoccyx audeberti a forest dependent species in West Africa? Malimbus, 16, 52–3. Alström, P. and Ranft, R. (2003). The use of sounds in bird systematics, and the importance of bird sound archives. In Why museums matter: avian archives in an age of extinction, (eds. N. Collar and C. Fisher), Bulletin of the British Ornithologists’ Club, Supplement, 123A, 114–35. Alvarez, F. (1994a). Cuckoo predation on nests of nearest neighbours of parasitized nests. Ardea, 82, 269–70. Alvarez, F. (1994b). Rates of weight increase of Cuckoo (Cuculus canorus) and hosts (Cercotrichas galactotes) chicks. Ardeola, 41, 63–5. Alvarez, F. (1994c). A gens of Cuckoo Cuculus canorus parasitizing Rufous Bush Chat Cercotrichas galactotes. Journal of Avian Biology, 25, 239–43. Alvarez, F. (1996). Model Cuckoo Cuculus canorus eggs accepted by Rufous Bush Chats Cercotrichas galactotes during the parasite’s absence from the breeding area. Ibis, 136, 340–2. Alvarez, F. and Arias de Reyna, L. (1974). Mecanismos de parasitación por Clamator glandarius y defensa por Pica pica. Doñana, Acta Vertebratica , 1, 43–65. Amadon, D. (1942). Birds collected during the Whitney South Sea Expedition. L. Notes on some non-passerine genera. American Museum Novitates, 1176. Amadon, D. (1953). Avian systematics and evolution in the Gulf of Guinea, the J. G. Correira collection. Bulletin of the American Museum of Natural History, 100, 393–452. Ambrose, S. J. and Murphy, D. P. (1994). Synchronous breeding of land birds on Barrow Island, Western Australia, after cyclonic summer rains. Emu, 94, 55–8.

Andersson, M. (1994). Sexual selection. Princeton University Press, Princeton. Andersson, M. (1995). Evolution of reversed sex roles, sexual size dimorphism, and mating system in coucals (Centropodidae, Aves). Biological Journal of the Linnean Society, 54, 173–81. Andersson, M. and Åhlund, M. (2000). Host-parasite relatedness shown by protein fingerprinting in a brood parasitic bird. Proceedings of the National Academy of Sciences USA, 97, 13188–93. Andersson, M. and Åhlund, M. (2001). Protein fingerprinting: a new technique reveals extensive conspecific brood parasitism. Ecology, 82, 1433–42. Andrew, P. (1990). The status of the Sunda Coucal Centropus nigrorufus Cuvier. Kukila, 5, 56–64. Andrew, P. (1992). The birds of Indonesia: a checklist (Peters’ sequence). Indonesian Ornithological Society, Jakarta. Andrew, P. and Holmes, D. A. (1990). Sulawesi bird report. Kukila, 5, 4–26. Andrle, R. F. (1967). Birds of the Sierra de Tuxtla in Veracruz, Mexico. Wilson Bulletin, 79, 163–87. Appanah, S. (1985). General flowering in the climax rain forests of south-east Asia. Journal of Tropical Ecology, 1, 225–40. Appert, O. (1967). Die Rachenzeichnung beim Nestling des Braunkopf-Seidenkuckucks Coua ruficeps olivaceiceps (Sharpe) von Madagaskar. Ornithologische Beobachter, 64, 52–6. Appert, O. (1970). Zur Biologie einiger Kua-Arten Madagaskars (Aves, Cuculi). Zoologische Jahrbücher.Abteilung für Systematik, Geographe und Biologie der Tiere , 97, 424–53. Appert, O. (1980). Erste Farbaufnahmen der Rachenzeichnung junger Kuas von Madagaskar (Cuculi, Couinae). Ornithologische Beobachter, 77, 85–101. Aragón, S., Møller, A. P., Soler, J. J., and Soler, M. (1999). Molecular phylogeny of cuckoos supports a polyphyletic

528 Bibliography origin of brood parasitism. Journal of Evolutionary Biology, 12, 495–506. Araya Mödinger, B. and Millie Holman, G. (1986). Guia de campo de las aves de Chile. Editorial Universitaria, Santiago.

Aveledo H., R. and Ginés, R. H. (1950). Descripción de cuatro aves nuevas de Venezuela. Sociedad de Ciencias Naturales la Salle Memoria, 10 (26), 59–71. Avise, J. C. (1994). Molecular markers, natural history and evolution. Chapman and Hall, New York.

Arbocco, G., Capocaccia, L., and Violani, C. (1978). Catalogo dei tipi di uccelli del Museo Civico di Storia Naturale di Genova. Annali del Museo Civico di Storia Naturale di Genova, 82, 184–265.

Aveledo Hostos, R. and Pérez Chinchilla, L. A. (1994). Descripcion de nueve subespecies nuevas y comentarios sobre dos especies de aves de Venezuela. Sociedad Venezolano de Ciencias Naturales Boletin, 44, 229–57.

Arias de Reyna, L. (1998). Coevolution of the Great Spotted Cuckoo and its hosts. In Parasitic birds and their hosts, (eds. S. I. Rothstein and S. K. Robinson), pp. 129–42. Oxford University Press, Oxford.

Azara, Felix de. (1809). Voyages dans l’Amérique méridionale, vol. 4. Dentu, Paris.

Arias de Reyna, L. and Hidalgo, S. J. (1982). An investigation into egg-acceptance by Azure-winged Magpies and host recognition by Great Spotted Cuckoo chicks. Animal Behaviour, 30, 819–23. Arias de Reyna, L., Recuerda, P., Corvillo, M., and Agular, I. (1982). Reproducción del críalo (Clamator glandarius) en Sierra Morena Central. Doñana, Acta Vertebrata, 9, 177–93. Arias de Reyna, L., Recuerda, P., Trujillo, J., Corvillo, M., and Cruz, A. (1987). Territory in the Great Spotted Cuckoo (Clamator glandarius). Journal für Ornithologie , 128, 231–9. Aristotle. (1991). History of animals, books VII–X, edited and translated by D. M. Balme. Loeb Classical Library, Harvard University Press, Cambridge, Massachusetts. Ash, J. S. and Miskell, J. E. (1998). Birds of Somalia. Pica Press, Sussex.

Baird, R. F. (1985a). Avian fossils from Quaternary deposits in ‘Green Waterhole Cave’, south-eastern South Australia. Records of the Australian Museum, 37, 353–70. Baird, R. F. (1985b). Centropus the giant coucal. In Kadimakara. Extinct vertebrates of Australia, (eds. P. V. Rich and G. F. van Tets), pp. 205–8. Pioneer Design Studio, Lilydale, Victoria. Baird, R. F. and Vickers-Rich, P. (1997). Eutreptodactylus itaboraiensis gen. et sp. nov., an early cuckoo (Aves: Cuculidae) from the Late Paleocene of Brazil. Alcheringa, 21, 123–7. Baker, E. C. S. (1906–1907). The oology of Indian parasitic cuckoos. Journal of the Bombay Natural History Society, 17, 72–84, 351–74, 678–96. Baker, E. C. S. (1907). Additional cuckoo notes. Journal of the Bombay Natural History Society, 17, 876–894. Baker, E. C. S. (1908a). Additional cuckoo notes. Journal of the Bombay Natural History Society, 18, 275–9. Baker, E. C. S. (1908b). The oology of parasitic cuckoos. Journal of the Bombay Natural History Society, 18, 915–6.

Aspinwall, D. R. and Beel, C. (1998). A field guide to Zambian birds not found in southern Africa. Zambian Ornithological Society, Lusaka.

Baker, E. C. S. (1919). Some notes on the genus Surniculus. Novitates Zoologicae, 26, 291–4.

Atkinson, R. W. (1982). Breeding the Renauld’s or Coralbilled Ground Cuckoo Carpococcyx renauldi at Metro Toronto Zoo. International Zoo Yearbook, 22, 168–71.

Baker, E. C. S. (1922). The fauna of British India including Ceylon and Burma. Birds, vol. 2, 1st edn. Taylor and Francis, London.

Audubon, J. J. (1839). Ornithological biography or an account of the habits of the birds of the United States of America, vol. 5. Adam and Charles Black, Edinburgh.

Baker, E. C. S. (1927). The fauna of British India including Ceylon and Burma. Birds, vol. 4, 2nd edn. Taylor and Francis, London.

Austin, O. L. (1948). The birds of Korea. Bulletin of the Museum of Comparative Zoology, Harvard 101.

Baker, E. C. S. (1934). The nidification of birds of the Indian Empire, vol. 3. Taylor and Francis, London.

Austin, O. L. and Kuroda, N. (1953). The birds of Japan, their status and distribution. Bulletin of the Museum of Comparative Zoology, Harvard 109, 277–637.

Baker, E. C. S. (1940). Cuculus canorus bakeri Hartert, Khasia Hills Cuckoo. In Life histories of North American cuckoos, goatsuckers, hummingbirds and their allies,(ed.

Bibliography 529 A. C. Bent), pp. 91–105. United States National Museum Bulletin 176.

Banks, E. (1935). Notes on birds in Sarawak, with a list of native names. Sarawak Museum Journal, 4, 267–325.

Ball, T. (2002). Birds of Thailand, songs and calls 2, cassette. Silkworm Books, Chiang Mai.

Banks, R. C. (1988a). Geographic variation in the Yellow-billed Cuckoo. Condor, 90, 473–7.

Baker, E. C. S. (1942). Cuckoo problems. H. F. and G. Witherby, London.

Banks, R. C. (1988b). An old record of the Pearlybreasted Cuckoo in North America and a nomenclatural critique. Bulletin of the British Ornithologists’ Club, 108, 87–91.

Bates, G. L. (1930). Handbook of the birds of West Africa. John Bale, Sons and Danielsson, London. Baker, N. E. and Baker, E. M. (1994). Dusky Long-tailed Cuckoo Cercococcyx mechowi and Papyrus Canary Serinus koliensis: two additions to the Tanzania list. Scopus, 18, 122–3. Baker, R. H. (1951). The avifauna of Micronesia, its origin, evolution and distribution. University of Kansas Publications Museum of Natural History, 3, 1–359. Baldamus, E. (1853). Neue Beiträge zur Fortpflanzungsgeschichte des europaischen Kukkuks, Cuculus canorus. pp. 307–25. Naumannia, Archiv fur die Ornithologie. Deutschen Ornithologische-Gesellschaft, Stuttgart. Baldamus, A. C. E. (1892). Das Leben der europäischen Kuckucke, nebst Beiträgen zur Lebenskunde der übringen parasitischen Kuckucke und Stärlinge . Paul Parey, Berlin. Bang, B. G. and Wenzel, B. M. (1985). Nasal cavity and olfactory system. In Form and function in birds, vol. 3, (eds. A. S. King and J. McLelland), pp. 195–225. Academic Press, New York. Bangs, O. (1918). Vertebrata from Madagascar. Bulletin of the Museum of Comparative Zoology, Harvard, 61, 87–511.

Banks, R. C. (1990). Geographic variation in the Yellow-billed Cuckoo: corrections and comments. Condor, 92, 538. Banks, R. C. and Hole, R., Jr. (1991). Taxonomic review of the Mangrove Cuckoo, Coccyzus minor (Gmelin). Caribbean Journal of Science, 27, 54–62. Bannerman, D. A. (1919). [Centropus efulensis neumanni Boyd Alexander]. Bulletin of the British Ornithologists’ Club, 39, 95–6. Bannerman, D. A. (1921). Remarks on rare and otherwise interesting birds contained in collections made by Mr. G. L. Bates in southern Cameroon. Ibis, Ser. 11, vol. 3, 81–121. Bannerman, D. A. (1933). The birds of tropical West Africa, vol. 3. Crown Agents, London. Bannerman, D. A. (1951). The birds of tropical West Africa, vol. 8. Crown Agents, London. Barbour, T. (1943). Cuban ornithology. Memoirs of the Nuttall Ornithological Club no. 9. Cambridge, Massachusetts.

Bangs, O. (1930). Types of birds now in the Museum of Comparative Zoölogy. Bulletin of the Museum of Comparative Zoology, Harvard, 70, 147–426.

Barlow, C. R. (2003). First conclusive evidence of breeding in Senegambia and parental behaviour of Black Coucal Centropus grillii. Bulletin of the African Bird Club, 10, 53–4.

Bangs, O. and Penard, T. E. (1918). Notes on a collection of Surinam birds. Bulletin of the Museum of Comparative Zoology, Harvard, 62, 23–93.

Barlow, C., Wacher, T., and Disley, T. (1997). A field guide to the birds of The Gambia and Senegal. Pica Press, Robertsbridge, U.K.

Bangs, O. and Penard, T. E. (1921). Notes on some American birds, chiefly Neotropical. Bulletin of the Museum of Comparative Zoology, Harvard, 64, 365–97.

Barnard, P. and Markus, M. B. (1990). Reproductive failure and nest site selection of two estrildid finches in Acacia woodland. Ostrich, 61, 117–24.

Bangs, O. and Phillips, J. C. (1914). Notes on a collection of birds from Yunnan. Bulletin of the Museum of Comparative Zoology, Harvard, 58, 267–302. Bangs, O. and Van Tyne, J. (1931). Birds of the KelleyRoosevelts Expedition to French Indo-China. Field Museum of Natural History Publication, 290, Zoology Series, 16, 31–119.

Barrett, C. L. (1905). Cuckoo notes. Emu, 5, 20–3. Barry, K. J. (1998). Interactions between a juvenile Striped Cuckoo and Arrow-marked Babblers. Honeyguide, 44, 89. Barton, R. A. (1996). Neocortex size and behavioural ecology in primates. Proceedings of the Royal Society of London, Series B, 263, 173–7.

530 Bibliography Barua, M. and Sharma, P. (1999). Birds of Kaziranga National Park, India. Forktail, 15, 47–60.

Becking, J. H. (1975b). New evidence of the specific affinity of Cuculus lepidus Müller. Ibis, 117, 275–84.

Bates, G. L. (1909). Field-notes on the birds of southern Kamerun, West Africa. Ibis, Ser. 9, vol. 3, 1–74.

Becking, J. H. (1981). Notes on the breeding of Indian cuckoos. Journal of the Bombay Natural History Society, 78, 201–31.

Bates, G. L. (1911). Further notes on the birds of southern Cameroon, part I. Ibis, Ser. 9, vol. 5, 479–545. Bates, G. L. (1924). On the birds collected in northwestern and northern Cameroon and parts of northern Nigeria. Ibis, Ser. 11, vol. 6, 1–45. Bates, G. L. (1927). Notes on some birds of Cameroon and the Lake Chad region: their status and breedingtimes. Ibis, Ser. 12, vol. 3, 1– 64 + 1 plate. Bates, G. L. (1930). Handbook of the birds of West Africa. John Bale, Sons and Danielsson, London. Bates, R. S. P. and Lowther, E. H. N. (1952). Breeding birds of Kashmir. Oxford University Press, Oxford. Baumel, J. J., ed. (1993). Handbook of avian anatomy: Nomina anatomica avium, 2nd edn. Publications of the Nuttall Ornithological Club, 23. Baumgartner, F. M. and Baumgartner, A. M. (1992). Oklahoma bird life. University of Oklahoma Press, Norman. Bayliss, M. J. (1989a). Cuckoo X breaks records. BTO News no. 159, 7. Bayliss, M. J. (1989b). Cuckoo egg-laying territories, from under productive co-existence to higher productive independence? BTO News no. 165, 11. Beal, F. E. L. (1898). Cuckoos and shrikes and their relation to agriculture. United States Department of Agriculture, Biological Survey Bulletin, 9. Beal, K. G. (1981). Winter foraging habits of the Roadrunner. Bulletin of the Oklahoma Ornithological Society, 14, 13–5. Beal, K. G. and Gillam, L. D. (1979). On the function of prey beating by Roadrunner. Condor, 81, 85–7. Beck, R. A., Burbank, D. W., Sercombe, W. J., Riley, G. W., Barndt, J. K., Berry, J. R., Afzal, J., Khan, A. M., Jurgen, H., Metje, J., Cheema, A., Shafique, N. A., Lawrence, R. D., and Khan, M. A. (1995). Stratigrapic evidence for an early collision between nortwest India and Asia. Nature, 373, 55–8. Becking, J. H. (1975a). The ultrastructure of the avian eggshell. Ibis, 117, 143–51.

Becking, J. H. (1988). The taxonomic status of the Madagascar Cuckoo Cuculus (poliocephalus) rochii and its occurrence on the African mainland, including southern Africa. Bulletin of the British Ornithologists’ Club, 108, 195–206. Beddard, F. E. (1885). On the structural characters and classification of the cuckoos. Proceedings of the Zoological Society of London, 1885, 168–87. Beddard, F. E. (1898a). On the anatomy of an Australian cuckoo, Scythrops novae-hollandiae. Proceedings of the Zoological Society of London, 1898, 44–9. Beddard, F. E. (1898b). The structure and classification of birds. Longman, London. Beddard, F. E. (1901). On the anatomy of the Radiated Fruit-cuckoo (Carpococcyx radiatus). Ibis, Ser. 8, vol. 1, 200–14. Beddard, F. E. (1902). On the syrinx and other points in the structure of Hierococcyx and some allied genera of cuckoos. Ibis, Ser. 8, vol. 2, 599–608. Beehler, B. M., Pratt, T. K., and Zimmerman, D. A. (1986). Birds of New Guinea. Princeton University Press, Princeton, New Jersey. Beisenherz, W. (1998). Mountain Tailorbird as host of Rusty-breasted Cuckoo. Kukila, 10, 159. Belcher, C. and Smooker, G. D. (1936). Birds of the colony of Trinidad and Tobago, part 3. Ibis, Ser. 13, vol. 6, 1–35. Bell, H. L. (1969). Recent Papuan breeding records. Emu, 69, 235–7. Bell, H. L. (1970). Field notes on the birds of Amazon Bay, Papua. Emu, 70, 23–6. Bell, H. L. (1979). The effect on rainforest birds of plantings of teak, Tectona grandis, in Papua New Guinea. Australian Wildlife Research, 6, 305–18. Bell, H. L. (1982a). A bird community of lowland rain forest in New Guinea. 1. Composition and density of avifauna. Emu, 82, 24–41. Bell, H. L. (1982b). A bird community of lowland rainforest in New Guinea. 2. Seasonality. Emu, 82, 65–74.

Bibliography 531 Bell, H. L. (1984). Carrying of young by coucals Centropus spp. Australian Bird Watcher, 10, 171. Belton, W. (1984). Birds of Rio Grande do Sul, Brazil. Part 1. Rheidae through Furnariidae. Bulletin of the American Museum of Natural History, 178, 371–631. Bencke, G. A. and Kindel, A. (1999). Bird counts along an altitudinal gradient of Atlantic forest in northeastern Rio Grande do Sul, Brazil. Ararajuba, 7, 91–107. Bendire, C. (1895). Life histories of North American birds. United States National Museum Special Bulletin no. 3. Benito-Espinal, E. (1990). Oiseaux des petites antilles. Editions du Latanier, Saint-Barthelemy. Bennett, A. G. (1937). Coccyzus melanocoryphus in the Falkland Islands. Ibis, Ser. 14, vol. 1, 868. Bennett, P. M. and Harvey, P. H. (1985). Relative brain size and ecology in birds. Journal of Zoology London, 207, 151– 69. Benson, C. W. (1942). Additional notes on Nyasaland birds. Ibis, 1942, 197–224. Benson, C. W. (1948). A new race of coucal from Nyasaland. Bulletin of the British Ornithologists’ Club, 68, 127–8. Benson, C. W. (1964). The species of Cercococcyx in Mwinilunga, Northern Rhodesia. Bulletin of the British Ornithologists’ Club, 84, 5–7. Benson, C. W. (1999). Type specimens of bird skins in the University Museum of Zoology, Cambridge, United Kingdom. British Ornithologists’ Club Occasional Publications, 4.

Benson, C. W. and Pitman, C. R. S. (1964). Further breeding records from Northern Rhodesia (No. 4). Bulletin of the British Ornithologists’ Club, 84, 54–69. Benson, C. W. and Schüz, E. (1971). A specimen of Coua delalandei (Temminck) (Cuculidae). Bulletin of the British Ornithologists’ Club, 91, 159–60. Benson, C. W., Brooke, R. K., Dowsett, R. J., and Irwin, M. P. S. (1970). Notes on the birds of Zambia: part V. Arnoldia (Rhodesia), 40 (4), 1–59. Benson, C. W., Brooke, R. K., Dowsett, R. J., and Irwin, M. P. S. (1971). The birds of Zambia. Collins, London. Benson, C. W., Colebrook-Robjent, J. F. R., and Williams, A. (1976). Contribution à l’ornithologie de Madagascar (suite). L’Oiseau et la Revue Française d’Ornithologie, 46, 209–42. Bent, A. C. (1940). Life histories of North American cuckoos, goatsuckers, hummingbirds, and their allies. United States National Museum Bulletin 176. Beresford, P. and Cracraft, J. (1999). Speciation in African forest robins (Stiphrornis): species limits, phylogenetic relationships, and molecular biogeography. American Museum Novitates, 3270. Berger, A. J. (1952). The comparative functional morphology of the pelvic appendage in three genera of Cuculidae. American Midland Naturalist, 47, 513–605. Berger, A. J. (1953a). On the locomotor anatomy of the Blue Coua, Coua caerulea. Auk, 70, 49–83. Berger, A. J. (1953b). The pterylosis of Coua caerulea. Wilson Bulletin, 65, 12–7.

Benson, C. W. and Benson, F. M. (1977). The birds of Malawi. Montfort Press, Limbe, Malawi.

Berger, A. J. (1960). Some anatomical characters of the Cuculidae and the Musophagidae. Wilson Bulletin, 72, 60–104.

Benson, C. W. and Irwin, M. P. S. (1965). The birds of Cryptosepalum forests, Zambia. Arnoldia (Rhodesia), 28, 1–12.

Berger, A. J. and Lunk, W. A. (1954). The pterylosis of the nestling Coua ruficeps. Wilson Bulletin, 66, 119–26. Berney, F. L. (1907). Cuckoos and crows. Emu, 7, 87.

Benson, C. W. and Irwin, M. P. S. (1972). The Thickbilled Cuckoo Pachycoccyx audeberti (Schlegel) (Aves: Cuculidae). Arnoldia (Rhodesia), 5 (33), 1–24.

Bernstein, H. A. (1859). Ueber Nester und Eier einiger javascher Vögel. Journal für Ornithologie , 7, 180–99.

Benson, C. W. and Irwin, M. P. S. (1973). Pachycoccyx audeberti: some addenda. Bulletin of the British Ornithologists’ Club, 93, 160–1.

Berthold, P., Fiedler, W., and Querner, U. (1995). Die Mönchsgrasmücke (Sylvia atricapilla) als Kuckucks (Cuculus canorus)-Wirt. Charadrius, 31, 11–8.

Benson, C. W. and Penny, M. J. (1971). The land birds of Aldabra. Philosophical Transactions of the Royal Society of London, Series B, 260, 417–527.

Bertram, B. C. R. (1970). The vocal behaviour of the Indian Hill Mynah, Gracula religiosa. Animal Behaviour Monographs, 3, 79–192.

532 Bibliography Beruldsen, G. (1980). A field guide to nests and eggs of Australian birds. Rigby, Adelaide. Beruldsen, G. (1990). Cape York in the wet. Australian Bird Watcher, 13, 209–17. Betts, F. N. (1947). Bird life in an Assam jungle. Journal of the Bombay Natural History Society, 46, 667–84. Betts, F. N. (1966). Notes on some resident breeding birds of southwest Kenya. Ibis, 108, 513–30.

Biswas, B. (1951). Notes on the taxonomic status of the Indian Plaintive Cuckoo Cuculus passerinus Vahl. Ibis, 93, 596–7. Biswas, B. (1960). The birds of Nepal, part 2. Journal of the Bombay Natural History Society, 57, 516–46. Blaber, S. J. M. (1990). A checklist and notes on the current status of the birds of New Georgia, Western Province, Solomon Islands. Emu, 90, 205–14.

Bharucha, E. K. (1982). Sunbirds fostering fledglings of the Plaintive Cuckoos. Journal of the Bombay Natural History Society, 79, 670–1.

Blaise, M. (1965). Contribution a l’étude de la reproduction du coucou gris Cuculus canorus dans la nord-est de la France. L’Oiseau et la Revue Française d’Ornithologie, 35, 87–116.

Biancucci, L. (1995). Una nueva especie para la avifauna Cordobesa y nuevas localidades para otras seis. El Hornero, 14, 76.

Blake, C. H. (1956). Some bird weights from Jamaica. Bird-Banding, 27, 174– 8.

Biddulph, C. H. (1956). Occurrence of the Redfaced Malkoha, Phoenicophaeus pyrrhocephalus (Pennant) in Madura District, Madras Presidency. Journal of the Bombay Natural History Society, 53, 697–8. Biddulph, J. (1881). The birds of Gilgit. Stray Feathers, 9, 301–66.

Blake, E. R. and Hanson, H. C. (1942). Notes on a collection of birds from Michoacan, Mexico. Field Museum of Natural History, Zoological Series, 22, 513–51. Blake, J. G., Loiselle, B. A., and Vande Weghe, J. P. (1990). Weights and measurements of some Central African birds. Gerfaut, 80, 3–11.

Binford, L. C. (1971). Roadrunner captures Orchard Oriole in California. California Birds, 2, 139.

Blakers, M., Davies, S. J. J. F., and Reilly, P. N. (1984). The atlas of Australian birds. Royal Australian Ornithologists Union, Melbourne, Victoria.

Binford, L. C. (1989). A distributional survey of the birds of the Mexican state of Oaxaca. Ornithological Monographs, 43.

Bleich, V. C. (1975). Roadrunner predation on ground squirrels in California. Auk, 92, 147–9.

Bingham, C. T. (1880). Additional notes on the birds of Tenasserim and specially on those of the Thoungyeen Valley. Stray Feathers, 9, 138–97. BirdLife International. (2001). Threatened birds of Asia: the BirdLife International Red Data Book. BirdLife International, Cambridge, U. K. Bishop, D. R., Bishop, D., and Bishop, M. S. (1995). Occurrence of Coppery Sunbird Nectarinia cuprea being parasitised by Klaas’s Cuckoo Chrysococcyx klaas. Babbler (Gaberone), 29–30 , 36. Bishop, K. D. (1982). Endemic birds of Biak Island. Report on the ICBP survey of Biak Island’s endemic avifauna. ICBP, Cambridge, U. K. Bishop, K. D. (1999). Preliminary notes on some birds in Bhutan. Forktail, 15, 87–91. Bishop, K. D. and Brickle, N. W. (1998). An annotated checklist of the birds of the Tanimbar Islands. Kukila, 10, 115–50.

Bluntschli, H. (1938). Ein eigenartiges Gaumenorgan beim Nestling des madagassischen Buschkuckucks Coua cristata L. Bio-Morphosis, 1, 265–72. Blyth, E. (1842). A monograph of the Indian and Malayan species of Cuculidae, or birds of the cuckoo family. Journal of the Asiatic Society, 12, 897–928. Blyth, E. (1845). Notes and descriptions of various new or little known species of birds. Journal of the Asiatic Society, 14, 198–204. Blyth, E. (1867). The ornithology of Ceylon. – a supplement to Dr. Jerdon’s ‘Birds of India’. Ibis, Ser. 2, vol. 3, 294–314. Bogert, C. (1937). The distribution and the migration of the Long-tailed Cuckoo (Urodynamis taitensis Sparrman). American Museum Novitates, 933, 1–12. Board, R. G. and Perrott, H. R. (1979). Vaterite, a constituent of the eggshells of the non-parasitic cuckoos, Guira guira and Crotophaga ani. Calcified Tissue International, 29, 63–9.

Bibliography 533 Boas, J. E. V. (1929). Biologisch-anatomische Studien über den Hals der Vögel. Kongelige Danske Videnskabernes Selskab Skrifter, Naturvidenskabernes mathemat. Afd., Ser. 9, 1, 105–222. Bock, W. U. and Miller, W. de W. (1959). The scansorial foot of the woodpeckers with comments on the evolution of perching and climbing feet in birds. American Museum Novitates, no. 1931. Boesman, P. (1998). Some new information on the distribution of Venezuelan birds. Continga, 9, 27–39. Boet, M. and Boet, M. (1996). Premier cas de nidification connu du coucou geai Clamator glandarius dans les Alpes-Maritimes. Alauda, 64, 370–1.

Bradley, P. (1985). Birds of the Cayman Islands. BOU Check-list No. 19. British Ornithologists’ Union, Tring. Bradley, D. and Wolff, T. (1958). The birds of Rennell Island. In The natural history of Rennell Island, British Solomon Islands, vol. 1., (ed. T. Wolff), pp. 85–120. Danish Science Press, Copenhagen. Braun, R. (1934). Biologische Notizen über einiger Vögel Nord-Angolas. Journal für Ornithologie , 82, 553–60. Bravery, J. A. (1967). Field notes on the Oriental Cuckoo. Emu, 66, 267–71. Bravery, J. A. (1970). The birds of Atherton Shire, Queensland. Emu, 70, 49–63.

Borges, S. H., Cohn-Haft, M., Carvalhaes, A. M. P., Henriques, L. M., Pacheco, J. F., and Whittaker, A. (2001). Birds of Jaú National Park, Brazilian Amazon: species check-list, biogeography and conservation. Ornitologia Neotropical, 12, 109–40.

Brazil, M. A. (1991). The birds of Japan. Smithsonian Institution Press, Washington, D.C.

Bosque, C. and Lentino, M. (1987). The passage of North American migratory land birds through xerophytic habitats on the western coast of Venezuela. Biotropica, 19, 267–73.

Briskie, J. V., Naugler, C. T., and Leech, S. M. (1994). Begging intensity of nestling birds varies with sibling relatedness. Proceedings of the Royal Society of London, Series B, 258, 73–8.

Bourns, F. S. and Worcester, D. C. (1894). Preliminary notes on the birds and mammals collected by the Menage Scientific Expedition to the Philippine Islands. Occasional Papers of the Minnesota Academy of Natural Sciences, 1, 1–64.

Britton, P. L. (1977). A Kenya record of Cercococcyx montanus at sea-level. Scopus, 1, 23–4.

Bowden, C. G. R. (2001). The birds of Mount Kupe, southwest Cameroon. Malimbus, 23, 13–44.

Britton, P. L. (1980b). A clarification of the status of the Red-chested Cuckoo Cuculus solitarius in lowland areas of East Africa. Scopus, 4, 71.

Bowdler, T. (ed.). (1861). The family Shakespeare. Longman, Green, Longman and Roberts, London. Bowen, B. S., Koford, R. R., and Vehrencamp, S. L. (1989). Dispersal in the communally breeding Groovebilled Ani (Crotophaga sulcirostris). Condor, 91, 52–64. Bowen, B. S. (2002). Groove-billed Ani Crotophaga sulcirostris. Birds of North America, 612, 1–15. Bowen, B. S., Koford, R. R., and Vehrencamp, S. L. (1991). Seasonal pattern of reverse mounting in the Groove-billed Ani (Crotophaga sulcirostris). Condor, 93, 159–63. Bowler, J. and Taylor, J. (1989). An annotated checklist of the birds of Manusela National Park, Seram (birds recorded on the Operation Raleigh Expedition). Kukila, 4, 3–29.

Brewster, W. (1902). Birds of the cape region of Lower California. Bulletin of the Museum of Comparative Zoology, Harvard, 41, 1–241.

Britton, P. L. (1980a). Birds of East Africa. East African Natural History Society, Nairobi.

Britton, P. L. (1982). Eurasian Cuckoo Cuculus canorus oversummering in East Africa. Scopus, 6, 45–6. Brodkorb, P. (1971). Catalogue of fossil birds: part 4 (Columbiformes through Piciformes). Bulletin of the Florida State Museum, Biological Sciences, 15 (4). Brooke, M. (1989). Tricks of the egg trade. Natural History, 89 (4), 50–4. Brooke, M. de L. (1995). The modern avifauna of the Pitcairn Islands. Biological Journal of the Linnean Society, 56, 199–212. Brooke, M. de L. and Davies, N. B. (1987). Recent changes in host usage by Cuckoos (Cuculus canorus) in Britain. Journal of Animal Ecology, 56, 873–83.

534 Bibliography Brooke, M. de L. and Davies, N. B. (1988). Egg mimicry by cuckoos Cuculus canorus in relation to discrimination by hosts. Nature, 335, 630–2. Brooke, M. de L. and Davies, N. B. (1989). Provisioning of nestling Cuckoos Cuculus canorus by Reed Warbler Acrocephalus scirpaceus hosts. Ibis, 131, 250–6.

Horsfield’s Bronze-cuckoo Chrysococcyx basalis and the Shining Bronze-cuckoo C. lucidus, in Western Australia: a new model for the evolution of egg morphology and host specificity in avian brood parasites. Ibis, 131, 528–47. Brooker, M. G. and Brooker, L. C. (1991). Eggshell strength in cuckoos and cowbirds. Ibis, 133, 406–13.

Brooke, M. de L. and Davies, N. B. (1991). A failure to demonstrate host imprinting in the Cuckoo Cuculus canorus and alternative hypotheses for the maintenance of egg mimicry. Ethology, 89, 154–66.

Brooker, M. G. and Brooker, L. C. (1992). Evidence for individual female host specificity in two Australian bronze-cuckoos (Chrysococcyx spp.). Australian Journal of Zoology, 40, 485–93.

Brooke, M. de L. and Nakamura, H. (1998). The acquisition of host-specific feather lice by Common Cuckoos (Cuculus canorus). Journal of Zoology, 244, 167–74.

Brooker, M. G. and Brooker, L. C. (1996). Acceptance by the Splendid Fairy-wren of parasitism by Horsfield’s Bronze-cuckoo: further evidence for evolutionary equilibrium in brood parasitism. Behavioral Ecology, 7, 395–407.

Brooke, M. de L., Davies, N. B., and Noble, D. G. (1998). Rapid decline of host defences in response to reduced Cuckoo parasitism: behavioural flexibility of Reed Warblers in a changing world. Proceedings of the Royal Society of London, Series B, 265, 1277–82. Brooke, R. K. (1964). Avian observations on a journey across central Africa and additional information on some of the species seen. Ostrich, 35, 277–92. Brooke, R. K. (1965a). Notes, chiefly distributional, on some birds in the Kafue National Park. Puku, 3, 59–65. Brooke, R. K. (1965b). Ornithological notes on the Furancungo district of Mozambique. Arnoldia Rhodesia, 10 (2), 1–10. Brooke, R. K., Crowe, T. M., and Chambal, M. S. (1990). Copulatory behaviour in Burchell’s Coucal Centropus burchellii. Ostrich, 61, 87. Brooker, L. C. and Brooker, M. G. (1990). Why are cuckoos host specific? Oikos, 57, 301–9. Brooker, L. C. and Brooker, M. G. (1994). A model for the effects of fire and fragmentation on the population viability of the Splendid Fairy-wren. Pacific Conservation Biology, 1, 344–58. Brooker, M. and Brooker, L. C. (1986). Identification and development of the nestling cuckoos, Chrysococcyx basalis and C. lucidus plagosus, in Western Australia. Australian Wildlife Research, 13, 197–202. Brooker, M. G. and Brooker, L. C. (1989a). Cuckoo hosts in Australia. Australian Zoological Reviews, No. 2, 1–67. Brooker, M. G. and Brooker, L. C. (1989b).The comparative breeding behaviour of two sympatric cuckoos,

Brooker, M. G., Brooker, L. C., and Rowley, I. (1988). Egg deposition by the bronze-cuckoos Chrysococcyx basalis and Ch. lucidus. Emu, 88, 107–9. Brooker, L. C., Brooker, M. G., and Brooker, A. M. H. (1990). An alternative population genetics model for the evolution of egg mimesis and egg crypsis in cuckoos. Journal of Theoretical Biology, 146, 123–43. Brooks, D. J., Evans, M. I., Martins, R. P., and Porter, R. F. (1987). The status of birds in North Yemen and the records of OSME Expedition in autumn 1985. Sandgrouse, 9, 4–66. Brooks, D. R. and McLennan, D. A. (1991). Phylogeny, ecology, and behavior. University of Chicago Press, Chicago. Brooks, T., Dutson, G., Gabutero, L., and Timmins, R. (1995). Siburan—key area for birds on Mindoro. Bulletin of the Oriental Bird Club, 21, 28–33. Brosset, A. (1976). Observations sur le parasitisme de la reproduction du coucou émeraude Chrysococcyx cupreus au Gabon. L’Oiseau et la Revue Française d’Ornithologie, 46, 201–8. Brosset, A. and Erard, C. (1986). Les oiseaux des régions forestières du nord-est du Gabon, vol. 1, Ecologie et comportement des espèces. Société Nationale de Protection de la Nature, Paris. Broughton, K. E., Middleton, A. L. A., and Bailey, E. D. (1987). Early vocalizations of the Brown-headed Cowbird and three host species. Bird Behaviour, 7, 27–30. Brown, C. R. (1984). Laying eggs in a neighbor’s nest: benefit and cost of colonial nesting in swallows. Science, 224, 518–9.

Bibliography 535 Brown, C. R. and Brown, M. B. (1988). A new form of reproductive parasitism in Cliff Swallows. Nature, London, 331, 66–8. Brown, C. R. and Brown, M. B. (1989). Behavioural dynamics of intraspecific brood parasitism in colonial Cliff Swallows. Animal Behaviour, 37, 777–96. Brown, C. R. and Brown, M. B. (1996). Coloniality in the Cliff Swallow, the effect of group size on social behavior. University of Chicago Press, Chicago. Brown, C. R. and Brown, M. B. (1998). Fitness components associated with alternative reproductive tactics in Cliff Swallows. Behavioral Ecology, 9, 158–71. Brown, J. L. (1987). Helping and communal breeding in birds: ecology and evolution. Princeton University Press, Princeton. Brown, L. H. and Britton, P. L. (1980). The breeding seasons of East African birds. East Africa Natural History Society, Nairobi. Brown, P. (1776). New illustrations of zoology. B. White, London. Browning, M. R. (1978). An evaluation of the new species and subspecies proposed in Oberholser’s Bird Life of Texas. Proceedings of the Biological Society of Washington, 91, 85–122. Bruce, M. D. (1987). Additions to the birds of Wallacea. 1. Bird records from smaller islands in the Lesser Sundas. Kukila, 3, 38–44. Bruggers, R. L. and Bortoli, L. (1979). Notes on breeding, parasitism and association with wasps of Heuglin’s Weaver nesting on telephone wires in Mali. Malimbus, 1, 135–44. Brunel, J. (1958). Observations sur les oiseaux du Bas-Dahomey. L’Oiseau et la Revue Française d’Ornithologie, 28, 1–38. Brunel, J. and Thiollay, J.-M. (1969). Liste préliminaire des oiseaux de Côte d’Ivoire. Alauda, 37, 230–54, 315–37. Brush, A. H. (1965). The structure and pigmentation of the feather tips of the Scaled Cuckoo. Auk, 82, 155–60.

Bryan, J. (1988). A jacobin summer. Oman Bird News, 3, 1–2. Bryant, H. C. (1916). Habits and food of the Roadrunner in California. University of California Publications in Zoology, 17, 21–58. Bryant, M. D. (1945). Phylogeny of Nearctic Sciuridae. American Midland Naturalist, 33, 257–390. Buckingham, R. and Jackson, L. (1990). A field guide to Australian birdsong. Cassette 5. Bird Observers Club of Australia, Nunawading, Victoria. Buden, D. W. (1987). The birds of the southern Bahamas. BOU Check-list No. 8, British Ornithologists’ Union, London. Buden, D. W. (1998). The birds of Kapingamarangi Atoll, including first record of the Shining Cuckoo (Chrysococcyx lucidus) from Micronesia. Notornis, 45, 151–3. Buden, D. W. (1999). The birds of Sapwuahfik Atoll, with first record of the Grey Wagtail, Motacilla cinerea, from the Federated States of Micronesia. Bulletin of the British Ornithologists’ Club, 119, 261–70. Burger, J. and Gochfeld, M. (2001). Smooth-billed Ani (Crotophaga ani) predation on butterflies in Mato Grosso, Brazil: risk decreases with increased group size. Behavioral Ecology and Sociobiology, 49, 482–92. Burney, D. A., James, H. F., Grady, F. V., Rafamantanantsoa, J.-G., Ramilisonina, M., Wright, H. T., and Cowart, J. B. (1997). Environmental change, extinction and human activity: evidence from caves in NW Madagascar. Journal of Biogeography, 24, 755–67. Bussmann, J. (1947). Wachstum und Jugendzeit eines Kuckucks. Ornithologische Beobachter, 44, 41–9. Butchert, S. H. M., Kilner, R. M., Fuisz, T., and Davies, N. B. (2003). Differences in the nestling begging calls of hosts and host-races of the Common Cuckoo, Cuculus canorus. Animal Behaviour, 65, 345–54.

Brush, A. H. (1979). Comparison of egg-white proteins: effect of electrophoretic conditions. Biochemical Systematics and Evolution, 7, 155–65.

Büttikofer, J. (1887). On a collection of birds made by Dr. C. Klaesi in the highlands of Padang (W. Sumatra) during the winter 1884–85. Notes from the Leyden Museum, 9, 1–96.

Brush, A. H. and Witt, H.-H. (1983). Intraordinal relationships of the Pelecaniformes and Cuculiformes: electrophoresis of feather keratins. Ibis, 125, 181–99.

Büttikofer, J. (1900). Zoological results of the Dutch Scientific Expedition to central Borneo, the birds. Notes from the Leyden Museum, 21, 145–276.

536 Bibliography Cabanis, J. and Heine, F. (1863). Museum Heineanum, Verzeichniss der ornithologischen Sammlung. 4 (1), Kuckucke und Faulvögel , 1–179. Cackett, K. (1970). Lowveld ornithological notes. Honeyguide, 1, 25–7. Cain, A. J. and Galbraith, I. C. J. (1956). Field notes on the birds of the eastern Solomon Islands. Ibis, 93, 100 –34, 262–95. Calder, W. A. (1967a). Breeding behavior of the Roadrunner, Geococcyx californianus. Auk, 84, 597–8. Calder, W. A. (1967b). Urine concentration of two carnivorous birds, the White Pelican and Roadrunner. Comparative Biochemistry and Physiology, 22, 607–9. Calder, W. A. (1968a). The diurnal activity of the Roadrunner, Geococcyx californianus. Condor, 70, 84–5. Calder, W. A. (1968b). Nest sanitation: a possible factor in the water economy of the Roadrunner. Condor, 70, 279. Calder, W. A. (1968c). There really is a Roadrunner. Natural History, 77 (4), 50–5. Calder, W. A. (1984). Size, function, and life history. Harvard University Press, Cambridge, Massachusetts.

Cariello, M. O., Lima, M. R., and Schwabl, H. G., and Macedo, R. H. F. (2004). Egg characteristics are unreliable in determining maternity in communal clutches of Guira Cuckoos Guira guira. Journal of Avian Biology, 35, 117–24. Carboneras, C. (1992). Family Anatidae (ducks, geese and swans). In Handbook of the birds of the world, volume 1, (eds. J. del Hoyo, A. Elliott, and J. Sargatal), pp. 536–628. BirdLife/Lynx Edicions, Barcelona. Carriker, M. A., Jr. and Meyer de Schauensee, R. (1935). An annotated list of two collections of Guatemalan birds in the Academy of Natural Sciences of Philadelphia. Proceedings of the Academy of Natural Sciences of Philadelphia, 87, 411–55. Cavalcanti, R. B., Lemes, M. R., and Cintra, R. (1991). Egg losses in communal nests of the Guira Cuckoo. Journal of Field Ornithology, 62, 177–80. Chalton, D. O. (1976). Weight loss in Spectacled Weavers raising a cuckoo chick. Ostrich, 47, 69. Chalton, D. O. (1991). Development of a Diederik Cuckoo chick in a Spectacled Weaver nest. Ostrich, 62, 84–5.

Calder, W. A. and Bentley, P. J. (1967). Urine concentration of two carnivorous birds, the White Pelican and Roadrunner. Comparative Biochemistry and Physiology A, 22, 607–9.

Chance, E. P. (1922). The Cuckoo’s secret. Sidgwick and Jackson, London.

Calder, W. A. and Schmidt-Nielsen, K. (1967). Temperature regulation and evaporation in the Pigeon and the Roadrunner. American Journal of Physiology, 213, 883–9.

Chandler, R. M. (1999). Fossil birds of Florissant, Colorado: with a description of a new genus and species of cuckoo. In National Park Service Paleontological Research, Technical Report NPS/NRGRD/GRDTR-99/03, vol. 4, (ed. V. L. Santucci and L. McClelland), pp. 49–53.

Campbell, A. J. (1900). Nests and eggs of Australian birds. A. J. Campbell, Melbourne. Capek, V. (1896). Beiträge zur Fortpflanzungsgeschichte des Kuckucks. Ornithologisches Jahrbuche, 7, 41–72, 102–17, 146–57, 165–83. Capper, D. R., Clay, R. P., Madroño N., A., Barnett, J. M., Burfield, I. J., Esquivel, E. Z., Kennedy, C. P., Perrens, M., and Pople, R. G. (2001). First records, noteworthy observations and new distributional data for birds in Paraguay. Bulletin of the British Ornithologists’ Club, 121, 23–37. Cariello, M. O., Schwabl, H. G., Lee, R. W., and Macedo, R. H. F. (2002). Individual female clutch identification through yolk protein electrophoresis in the communally breeding Guira Cuckoo. Molecular Ecology, 11, 2417–24.

Chance, E. P. (1940). The truth about the Cuckoo. Country Life Ltd., London.

Chapin, J. P. (1928). The African cuckoos of the genus Cercococcyx. American Museum Novitates, 313, 1–11. Chapin, J. P. (1939). The birds of the Belgian Congo, part II. Bulletin of the American Museum of Natural History, 75. Chappuis, C. (1974). Les oiseaux de l’ouest africain. Disk 1: Columbidae et Cuculidae. Alauda sound suppl. 1, Paris (Alauda, 42, 197–222, accompanying commentary). Chappuis, C. (2000). Oiseaux d’Afrique/African bird sounds. Alauda/ Société Ornithologique de France, Paris. set of 15 compact disks with accompanying text. Charles-Dominique, P. (1976). Les gommes dans le regime alimentaire de Coua cristata à Madagascar.

Bibliography 537 L’Oiseau et la Revue Française d’Ornithologie, 46, 174–8. Chasen, F. N. (1939). The birds of the Malay Peninsula, vol. 4. H. F. and G. Witherby, London. Chasen, F. N. and Kloss, C. B. (1927). Notes on Oriental birds. Journal of the Federated Malay States Museums, 13, 275–80. Chasen, F. N. and Kloss, C. B. (1930). On a collection of birds from the lowlands and islands of North Borneo. Bulletin of the Raffles Museum, 4, 1–12. Chavigny, J. de and Le Dû, R. (1938). Note sur l’adaptation des oeufs du coucou de l’Afrique du Nord Cuculus canorus bangsi suivie de quelques observations biologiques. Alauda, 10, 91–115. Cheke, R. A. and Walsh, J. F. (1996). The birds of Togo. BOU Check-list No. 14, British Ornithologists’ Union, London. Cheng Tso-Hsin (Zheng Zuoxin). (1991). Fauna Sinica, Aves, vol. 6. Columbiformes, Psittaciformes, Cuculiformes and Strigiformes. Science Press, Beijing. Cherrie, G. K. (1916). A contribution to the ornithology of the Orinoco region. Museum of the Brooklyn Institute of Arts and Sciences, Science Bulletin, 2, 133a–374. Chesser, R. T. (1994). Migration in South America: an overview of the austral system. Bird Conservation International, 4, 91–107. Chesser, R. T. (1997). Patterns of seasonal and geographical distribution of austral migrant flycatchers (Tyrannidae) in Bolivia. Ornithological Monographs, 48, 171–204. Chisholm, A. H. (1973). Cuckoos are very resolute. Australian Bird Watcher, 5, 49–54. Chittenden, H. (1995). Nocturnal calling of Barred Cuckoo. Birding in Southern Africa, 47, 49. Christison, P., Buxton, A., and Emmet, A. M., assisted by Ripley, D. (1946). Field notes on the birds of coastal Arakan and the foothills of the Yomas. Journal of the Bombay Natural History Society, 46, 13–32. Christy, P. and Clarke, W. (1994). Guide des oiseaux de la Réserve de la Lopé. ECOFAC Gabon, Libreville. Christy, P. and Clarke, W. V. (1998). Guide des oiseaux de São Tomé et Príncipe . ECOFAC, São Tomé. Chubb, C. (1916). The birds of British Guiana, Vol. 1. Bernard Quaritch, London.

Cibois, A., Pasquet, E., and Schulenberg, T. S. (1999). Molecular systematics of the Malagasy babblers (Passeriformes: Timaliidae) and warblers (Passeriformes: Sylviidae), based on cytochrome b and 16S rRNA sequences. Molecular Phylogenetics and Evolution, 13, 581–95. Cibois, A., Slikas, B., Schulenberg, T. S., and Pasquet, E. (2001). An endemic radiation of Malagasy songbirds is revealed by mitochondrial DNA sequence data. Evolution, 55, 1198–206. Cintra, R. and Sanaiotti, T. M. (1990). Osteocephalus taurinus (treefrog): predation. Herpetological Review, 21, 59. Claffey, P. (1998). The status of Black Cuckoo Cuculus clamosus and Red-chested Cuckoo C. solitarius in Benin. Malimbus, 20, 56–7. Clancey, P. A. (1951). Notes on birds of the South African subcontinent. Annals of the Natal Museum, 12, 137–52. Clancey, P. A. (1960). Miscellaneous taxonomic notes on African birds, 15. The races of the Jacobin Cuckoo Clamator jacobinus (Boddaert) occurring in Africa south of the Zambesi River. Durban Museum Novitates, 6 (2), 27–31. Clancey, P. A. (1961). The eastern cuckoo Cuculus canorus telephonus Heine as a South African bird. Ostrich, 32, 48–50. Clancey, P. A. (1962). Miscellaneous taxonomic notes on African birds XIX. Durban Museum Novitates, 6, 181–94. Clancey, P. A. (1980). SAOS checklist of southern African birds. Southern African Ornithological Society, Pretoria. Clancey, P. A. (1989). Four additional species of southern African endemic birds. Durban Museum Novitates, 14, 148–9. Clancey, P. A. (1990). Size-variation and post-breeding movement in the Didric Cuckoo Chrysococcyx caprius (Boddaert). Bulletin of the British Ornithologists’ Club, 110, 130–7. Clancey, P. A. (1996). The birds of southern Mozambique. African Bird Book Publishing, Westville. Clay, R. P., Capper, D. R., Barnett, J. M., Burfield, I. J., Esquivel, E. Z., Fariña, R., Kennedy, C. P., Perrens, M., and Pople, R. G. (1998). White-winged Nightjars Caprimulgus candicans and cerrado conservation: the key findings of Project Aguará Ñu 1997. Continga, 9, 52–6.

538 Bibliography Clayton, N. S., Reboreda, J. C., and Kacelnik, A. (1997). Seasonal changes of hippocampus volume in parasitic cowbirds. Behavioural Processes, 41, 237–43.

Collar, N. J. and Long, A. J. (1996). Taxonomy and names of Carpococcyx cuckoos from the Greater Sundas. Forktail, 11, 135–50.

Clayton, N. S. (1998). Memory and the hippocampus in food-storing birds: a comparative approach. Neuropharmacology, 37, 441–52.

Collar, N. J. and Stuart, S. N. (1985). Threatened birds of Africa and related islands,The ICBP/IUCN red data book, Part 1. (3rd edn.) International Council for Bird Preservation, Cambridge, U. K.

Clements, J. F. and Shany, N. (2001). A field guide to the birds of Peru. Ibis Publishing Company, Temecula, California. Clunie, F. (1973). Fan-tailed Cuckoo parasitizes Fiji Warbler. Notornis, 20, 168. Clunie, F. and Morse, P. (1984). Birds of the Fiji bush. Fiji Museum, Suva. Clutton-Brock, T. H. (1991). The evolution of parental care. Princeton University Press, Princeton. Coates, B. J. (1985). The birds of Papua New Guinea, vol. 1, Non-Passerines. Dove, Alderley, Queensland. Coates, B. J. (2001). Birds of New Guinea and the Bismarck Archipelago: a photographic guide. Dove, Alderley, Queensland.

Collar, N. J., Crosby, M. J., and Stuttersfield, A. J. (1994). Birds to watch 2: the world list of threatened birds. BirdLife International, Cambridge, U. K. Collias, N. E. and Collias, E. C. (1959). Breeding behaviour of the Black-headed Weaverbird Textor cucullatus graueri (Hartert) in the Belgian Congo. Ostrich, Supplement, 3, 233–41. Collias, N. E. and Collias, E. C. (1970). The behaviour of the West African Village Weaverbird. Ibis, 112, 457–80. Collins, N. M., Sayer, J. A., and Whitmore, T. C, eds. (1991). The conservation atlas of tropical forests: Asia and the Pacific. Macmillan, London.

Coates, B. J. and Bishop, K. D. (1997). A guide to the birds of Wallacea (Sulawesi, the Moluccas and Lesser Sunda Islands, Indonesia). Dove, Alderley, Queensland.

Collona, M., Casanova, J., Dullo, W.-C., and Camoin, G. (1996). Sea-level changes and δ18O record for the past 34,000 yr from Mayotte Reef, Indian Ocean. Quaternary Research, 46, 335–9.

Cockburn, A. (1998). Evolution of helping behavior in cooperatively breeding birds. Annual Review of Ecology and Systematics, 29, 141–77.

Colston, P. R. and Curry-Lindahl, K. (1986). The birds of Mount Nimba, Liberia. London: British Museum (Natural History).

Cocker, C. and Inskipp, C. (1988). A Himalayan ornithologist. The life and work of Brian Houghton Hodgson. Oxford University Press, Oxford.

Cometti, R. (1986). Little Barrier Island, New Zealand’s foremost wildlife sanctuary. Hodder and Stroughton, Auckland.

Cohn-Haft, M., Whittaker, A., and Stouffer, P. C. (1997). A new look at the “species-poor” central Amazon: the avifauna north of Manaus, Brazil. Ornithological Monographs, 48, 205–35.

Compton, S. G., Craig, A. J. F. K., and Walters, I. W. R. (1996). Seed dispersal in an African fig tree: bird as high quantity, low quality dispersers? Journal of Biogeography, 23, 553–63.

Cole, C. F. (1908). Observations on the finches as foster parents to the cuckoos. Emu, 8, 23–5.

Condon, H. T. (1947). A note on the calls of the Pallid Cuckoo. South Australian Ornithologist, 18, 46.

Colebrook-Robjent, J. F. R. (1977). Cuckoos in Zambia. Black Lechwe, 12 (3), 26–31.

Connop, S. (1993). Birdsongs of Nepal. Library of Natural Sounds, Cornell Laboratory of Ornithology, Ithaca.

Colebrook-Robjent, J. F. R. (1978). An oviduct egg of the Indian Cuckoo Cuculus micropterus. Bulletin of the British Ornithologists’ Club, 98, 40.

Connop, S. (1995). Birdsongs of the Himalayas. Turaco Nature Service, Toronto.

Colebrook-Robjent, J. F. R. (1984). The breeding of the Didric Cuckoo Chrysococcyx caprius in Zambia. Proceedings of the V Pan-African Ornithological Congress: 763–78.

Contreras, A. O. (1993). Hallazgo del yasy yatere guazu Dromococcyx phasianellus phasianellus (Spix, 1824) en isla Hu, Neembucu, sudeste del Paraguay oriental (Aves: Cuculidae, Neomorphinae). Notulas Faunisticas, 41, 1–2.

Bibliography 539 Contreras, J. R. (1979). Bird weights from northeastern Argentina. Bulletin of the British Ornithologists’ Club, 99, 21–4. Cooke, F. (1987). Lesser Snow Goose: a long-term population study. In Avian genetics, (eds. F. Cooke and P. A. Buckley), pp. 407–32. Academic Press, New York. Coomans de Ruiter, L. (1947). Nest and eggs of the Rufous-bellied Malkoha, Rhopodytes sumatranus (Raffles) in Dutch West Borneo. Limosa, 20, 187. Coomans de Ruiter, L. (1950). Vogel van het Quarles-Gebergte (ZW. Centraal Celebes). Ardea, 38, 40–182. Coomans de Ruiter, L. and Maurenbrecher, L. L. A. (1948). Stadsvogels van Makassar (Zuid-Celebes). Ardea, 36, 163–98. Cooper, J. H. (2000). First fossil record of Azure-winged Magpie Cyanxopica cyanus in Europe. Ibis, 142, 150–1. Cooper, R. P. (1953). Pheasant-Coucal. Australian Birds, 83. Cooper, R. P. (1958). Pallid Cuckoo feeding young. Emu, 58, 6 7–8. Cornwallis, L. and Porter, R. F. (1982). Spring observations on the birds of North Yemen. Sandgrouse, 4, 1–36. Cory, C. B. (1919). Catalogue of birds of the Americas, part 2, no. 1. Field Museum of Natural History Publication 197, Zoology Series, 13. Courtney, J. and Marchant, S. (1971). Breeding details of some common birds in south-eastern Australia. Emu, 71, 121–33. Cracraft, J. L. (1964). Observations on sun-bathing in the Yellow-billed Cuckoo. Wilson Bulletin, 76, 187.

Cracraft, J. (1988). From Malaysia to New Guinea: evolutionary biogeography within a complex continent-island arc contact zone. Acta XIX Congressus Internationalis Ornithologici, vol. 2, Ottawa, 2581–93. Cracraft, J. (2001). Avian evolution, Gondwana biogeography and the Cretaceous-Tertiary mass extinction event. Proceedings of the Royal Society of London, Series B, 268, 459–69. Cracraft, J. and Prum, R. O. (1988). Patterns and processes of diversification: speciation and historical congruence in some Neotropical birds. Evolution, 42, 603–20. Craig, A. J. F. K. and Hartley, A. H. (1985). The arrangement and structure of feather melanin granules as a taxonomic character in African starlings (Sturnidae). Auk, 102, 629–32. Cramp, S., ed. (1985). The birds of the western Palearctic, vol. 4. Oxford University Press, Oxford. Cramp, S., et al., eds. (1994). The birds of the western Palearctic, vol. 8. Oxford University Press, Oxford. Cranbrook, E. and Wells, D. R. (1981). Observations of fledgling cuckoos and their fosters in Gunung Mulu National Park. Sarawak Museum Journal, 29, 147–9. Crawford, D. N. (1972). Birds of Darwin area, with some records from other parts of Northern Territory. Emu, 72, 131–48. Crawford, R. L. and Stevenson, H. M. (1984). Patterns of spring and fall migration in northwest Florida. Journal of Field Ornithology, 55, 196–203. Crisler, T., Jameson, C., and Brouwer, J. (2003). An updated overview of the birds of W National Park, southwest Niger. Malimbus, 25, 4–30.

Cracraft, J. (1974). Phylogeny and evolution of the ratite birds. Ibis, 116, 494–520.

Crouch, H. and Crouch, A. (1993). Papua New Guinea bird calls. Papua New Guinea Bird Society, two cassettes, Boroko, PNG.

Cracraft, J. (1982). Geographic differentiation, cladistics, and vicariance biogeography: reconstructing the tempo and mode of evolution. American Zoologist, 22, 411–24.

Crouther, M. M. (1980). Some further observations of Channel-billed Cuckoos in North Queensland. Bird Observer, 583, 60–1.

Cracraft, J. (1983). Species concepts and speciation analysis. Current Ornithology, 1, 159–87.

Crouther, M. M. (1985). Some breeding records of the Common Koel Eudynamis scolopacea. Australian Bird Watcher, 11, 49–56.

Cracraft, J. (1985). Historical biogeography and patterns of differentiation within the South American avifauna. Ornithological Monographs, 36, 49–84.

Crouther, M. M. and Crouther, M. J. (1984). Observations on the breeding of Figbirds and Common Koels. Corella, 8, 89–92.

540 Bibliography Cruz, A. (1973). Food and foraging ecology of the Chestnut-bellied Cuckoo. Wilson Bulletin, 85, 336–7. Cruz, A. (1975). Ecology and behavior of the Jamaican Lizard Cuckoo. Studies on the Fauna of Curaçao and other Caribbean islands, 46, 109–11.

seconda e terza esplorazione del Fiume Fly negli anni 1876 e 1877. Annali del Museo Civico di Storia Naturale di Genova, 14, 21–147. Dammerman, K. W. (1922). The fauna of Krakatau, Verlaten Island and Sebesy. Treubia, 3, 61–112.

Cruz, A. and Wiley, J. W. (1989). The decline of an adaptation in the absence of a presumed selection pressure. Evolution, 43, 55–62.

Danforth, S. T. (1925). Birds of the Cartagena Lagoon, Porto Rico. Journal of the Department of Agriculture of Porto Rico, 10, 1–136.

Cuadros, T. (1991). Registro visual del cuco terrestre piquirrojo (Neomorphus pucheranii) en Colombia. Boletín SAO: Organo de publicación de la Sociedad Antioqueña de Ornitología, 2, 26–7.

Danforth, S. T. (1928). Bird observations in Jamaica during the summer of 1926. Auk, 45, 480–91.

Cuello, J. and Gerzenstein, E. (1962). Las aves del Uruguay, lista sistematica, distribucion y totas. Comunicaciones Zoologicas del Museo de Historia Natural de Montevideo, 6, 1–190. Curio, E. (1970). Eine Begattung des Eidechsenkuckucks (Saurothera vetula) auf Jamaica. Journal für Ornithologie , 111, 240. Curio, E., Hornbuckle, Y. de S., Aston, P., and Lastimoza, L. L. (2001). New bird records for the island of Panay, Philippines, including the first record of the Asian Stubtail Urosphena squamiceips for the Philippines. Bulletin of the British Ornithologists’ Club, 121, 183–97. Curran, L. M. and Leighton, M. (2000). Vertebrate responses to spatiotemporal variation in seed production of mast-fruiting Dipterocarpaceae. Ecological Monographs, 70, 101–28. Curran, L. M., Caniago, I., Paoli, G. D., Astianti, D., Kusneti, M., Leighton, M., Nirarita, C. E., and Haeruman, H. (1999). Impact of El Niño and logging on canopy tree recruitment in Borneo. Science, 286, 2184–8. Cyrus, D. and Robson, N. (1980). Bird atlas of Natal. University of Natal, Pietermaritzburg. Dabbene, R. (1926). Nidificacion del cuculido Coccyzus melacoryphus Vieill. Hornero, 3, 248–52. Daguerre, J. B. (1924). Apuntos sobre algunas aves de la provincia de Buenos Aires. Hornero, 3, 248–52. Dahl, F. (1899). Das Leben der Vögel auf den Bismarckinseln. Mitteilungen aus dem Zoologischen Museum in Berlin, 1, 108–222. D’Albertis, L. M. and Salvadori, T. (1879). Catalogo degli uccelli raccolti da L. M. D’Albertis durante la

Danilowitsch, A. P. (1934). Zur Verbreitung des Cuculus optatus Gould im europäischen Russland. Ornithologische Monatsberichte, 42, 155–6. Darwin, C. (1859). On the origin of species. John Murray, London. Darwin, C. (1875). On the origin of species, 6th ed. John Murray, London. David, A. and Oustalet, M. E. (1877). Les Oiseaux de la Chine. Vols. 1, 2. G. Masson, Ed. Libraire de l’Academie de Médicine, Paris. Davidar, P., Yoganand, T. R. K., Ganesh, T., and Joshi, N. (1996). An assessment of common and rare forest bird species of the Andaman Islands. Forktail, 12, 135–42. Davidson, J. (1886). Bird nesting on the Ghats. Journal of the Bombay Natural History Society, 1, 175–83. Davies, N. B. (2000). Cuckoos, cowbirds and other cheats. A. D. Poyser, London. Davies, N. B. and Brooke, M. de L. (1988). Cuckoos versus Reed Warblers: adaptations and counteradaptations. Animal Behaviour, 36, 262–84. Davies, N. B. and Brooke, M. de L. (1989a). An experimental study of co-evolution between the Cuckoo, Cuculus canorus, and its hosts. I. Host egg discrimination. Journal of Animal Ecology, 58, 207–24. Davies, N. B. and Brooke, M. de L. (1989b). An experimental study of co-evolution between the Cuckoo, Cuculus canorus, and its hosts. II. Host egg markings, chick discrimination and general discussion. Journal of Animal Ecology, 58, 225–36. Davies, N. B., Bourke, A. F. G., and Brooke, M. de L. (1989). Cuckoos and parasitic ants: interspecific brood parasitism as an evolutionary arms race. Trends in Ecology and Evolution, 4, 274–8.

Bibliography 541 Davies, N. B., Brooke, M. de L., and Kacelnik, A. (1996). Recognition errors and probability of parasitism determine whether Reed Warblers should accept or reject Cuckoo eggs. Proceedings of the Royal Society of London, Series B, 263, 925–31. Davies, N. B., Kilner, R. M., and Noble, D. G. (1998). Nestling cuckoos, Cuculus canorus, exploit hosts with begging calls that mimic a brood. Proceedings of the Royal Society of London, Series B, 263, 673–8. Davis, D. E. (1940a). Social nesting habits of the Smooth-billed Ani. Auk, 57, 179–218. Davis, D. E. (1940b). Social nesting habits of Guira guira. Auk, 57, 472–84. Davis, D. E. (1941). Social nesting habits of Crotophaga major. Auk, 58, 179–83. Davis, D. E. (1942). The phylogeny of social nesting habits in the Crotophaginae. Quarterly Review of Biology, 17, 115–34. Davis, J. (1960). Notes on the birds of Colima, Mexico. Condor, 62, 215–9. Davis, S. E. (1993). Seasonal status, relative abundance, and behavior of the birds of Concepción, Departamento Santa Cruz, Bolivia. Fieldiana Zoology, New Series 71, Publ. 1444, 1–33. Davison, G. W. H. (1979). A survey of terrestrial birds in the Gunung Mulu National Park, Sarawak. Sarawak Museum Journal, 27, 283–93. Davison, G. W. H. (1997). Bird observations in the Muratus Mountains, Kalimantan Selatan. Kukila, 9, 114–21. Davison, G. W. H. and Fook, Chew Yen. (1995). A photographic guide to birds of peninsular Malaysia and Singapore. New Holland Publishers, London. Davison, G. W. H. and Fook, Chew Yen. (1996). A photographic guide to birds of Borneo. New Holland Publishers, London. Davison, G. W. H. (1979). A survey of terrestrial birds in the Gunung Mulu National Park, Sarawak. Sarawak Museum Journal, 27, 283–93. Dean, W. R. J. (1971). Breeding data for the birds of Natal and Zululand. Durban Museum Novitates, 9, 59–91. Dean, W. R. J. (2000). The birds of Angola. BOU Checklist No. 18. British Ornithologists’ Union, Tring.

Dean, W. R. J., Macdonald, I. A. W., and Vernon, C. J. (1974). Possible breeding record of Cercococcyx montanus. Ostrich, 45, 188. Dean, W. R. J., Huntley, M. A., Huntley, B. J., and Vernon, C. J. (1988). Notes on some birds of Angola. Durban Museum Novitates, 14, 43–92. Dearborn, D. C. (1998). Begging behavior and food acquisition by Brown-headed Cowbird nestlings. Behavioral Ecology and Sociobiology, 43, 259–70. Dearborn, D. C. (1999). Brown-headed Cowbird nestling vocalizations and risk of nest predation. Auk, 116, 448–457. Debout, G., Meister, P., and Ventelon, M. (2000). Notes complémentaires sur l’avifaune du Niger. Malimbus, 22, 87–99. Deignan, H. G. (1945). The birds of northern Thailand. United States National Museum Bulletin, 186. Deignan, H. (1952). The correct name for the MalayoSumatran race of the Chestnut-breasted Malkoha (Cuculidae). Bulletin of the Raffles Museum, 24, 219. Deignan, H. G. (1955). Four new races of birds from East Asia. Proceedings of the Biological Society of Washington, 68, 145–58. Deignan, H. G. (1963). Checklist of the birds of Thailand. United States National Museum Bulletin, 226. Deignan, H. G. (1965). Type specimens of birds in the United States National Museum. United States National Museum Bulletin, 221. Deignan, H. G. and Amos, B. (1950). Notes on some forms of the genus Chalcites Lesson. Emu, 49, 167–8. Delacour, J. (1927). Renauld’s Ground Cuckoo (Carpococcyx renauldi). Avicultural Magazine, 4, 229–30. Delacour, J. (1932). Les oiseaux de la Mission FrancoAnglo-Américaine à Madagascar. L’Oiseau et la Revue Française d’Ornithologie, 2, 1–96. Delacour, J. (1947). Birds of Malaysia. Macmillan, New York. Delacour, J. and Jabouille, P. (1931). Les oiseaux de l’Indochine Française, vol. 2. Exposition Coloniale Internationale, Paris. Delacour, J. T. and Mayr, E. (1945). Notes on the taxonomy of the birds of the Philippines. Zoologica, 30, 105–17.

542 Bibliography Delacour, J. and Mayr, E. (1946). Birds of the Philippines. Macmillan, New York. Dement’ev, G. P. and Gladkov, N. A., eds. (1966). Birds of the Soviet Union, vol. 1. Israel Program for Scientific Translations, Jerusalem. Demey, R., Dowsett, R. J., and Fishpool, L. D. C. (2001). Comments on Black-throated Coucal Centropus leucogaster, claimed from Niger. Malimbus, 23, 112–113.

Diamond, J. M., Gilpin, M. E., and Mayr, E. (1976). Species-distance relation for birds of the Solomon Archipelago, and the paradox of the great speciators. Proceedings of the National Academy of Sciences USA, 73, 2160–4. Dias, M., Asensio, B., and Tellería, J. L. (1996). Aves Ibéricas, vol. 1, No Passeriformes. J. M. Reyero, Madrid.

Desfayes, M. (1974). More cuckoo problems. Journal of the Bombay Natural History Society, 71, 145–6.

Di Carlo, E. (1971). Appunti sulla biologia del cuculo cal ciuffo (Clamator glandarius). Rivista Italiana di Ornitologia, 41, 86–107.

Desjardins, P. and Morais, R. (1990). Sequence and gene organization of the chicken mitochondrial genome. Journal of Molecular Biology, 212, 599–634.

Dickerman, R. W. (1994). Notes on birds from Africa with descriptions of three new subspecies. Bulletin of the British Ornithologists’ Club, 114, 274–8.

Dewar, D. (1908). An enquiry into the parasitic habits of the Indian Koel. Journal of the Bombay Natural History Society, 17, 765–82.

Dickey, D. R. and van Rossem, A. J. (1938). The birds of El Salvador. Field Museum of Natural History, Zoology Series, Publication, 23 (406).

Dharmakumarsinhji, R. S. (1954). Birds of Saurashtra. The author, Bombay.

Dickinson, E. C. and Heucke, J. (1986). Notes on Philippine birds, 8. A collection from Mindoro revisited. Bulletin of the British Ornithologists’ Club, 106, 56–63.

Dhindsa, M. S. (1983). Intraspecific nest parasitism in two species of Indian weaverbirds Ploceus benghalensis and P. manyar. Ibis, 125, 243–5. Dhindsa, M. S. (1990). Intraspecific brood parasitism in the Baya Weaverbird Ploceus philippinus. Bird Behaviour, 8, 111–3. Dhindsa, M. S. and Toor, H. S. (1981). Some observations on a nest of the Common Crow-pheasant, Centropus sinensis (Stephens). Journal of the Bombay Natural History Society, 78, 600–2. Diamond, J. M. (1972). Avifauna of the eastern highlands of New Guinea. Publications of the Nuttall Ornithological Club, 12. Diamond, J. M. (1975). Distributional ecology and habits of some Bougainville birds (Solomon Islands). Condor, 77, 14–23. Diamond, J. M. (2002). Dispersal, mimicry, and geographic variation in northern Melanesian birds. Pacific Science, 56, 1–22. Diamond, J. M. and LeCroy, M. (1979). Birds of Karkar and Bagabag Islands, New Guinea. Bulletin of the American Museum of Natural History, 164, 467–531. Diamond, J. M. and Marshall, A. G. (1977). Distributional ecology of New Hebridean birds: a species kaleidoscope. Journal of Animal Ecology, 46, 703–27.

Dickinson, E. C., Kennedy, R. S., and Parkes, K. C. (1991). The birds of the Philippines. BOU Check-list No. 12, British Ornithologists’ Union, London. Dique, W. F. (1880). Letter to the editor. Stray Feathers, 9, 298. Dobie, J. F. (1939). The Roadrunner in fact and folklore. In In the shadow of history (eds. J. F. Dobie, M. C. Boatright and H. H. Ransom), pp. 146–174. Texas FolkLore Society, Austin. Dolan, B. (1939). Zoological results of the Second Dolan Expedition to Western China and Eastern Tibet, 1934–1936. Proceedings of the Academy of Natural Sciences of Philadelphia, 90, 159–84. Dorst, J. (1951). Contribution a l’étude du plumage des coucous métalliques du genre Chrysococcyx Boie. Bulletin du Muséum National d’Histoire Naturelle, 2nd série , 23, 173–80. Doughty, C., Day, N., and Plant, A. (1999). Birds of the Solomons,Vanuatu and New Caledonia. Christopher Helm, London. Dow, D. D. (1972). The New Zealand Long-tailed Cuckoo: nest parasite or predator? Emu, 72, 179–80. Downer, A. and Sutton, R. (1990). Birds of Jamaica. Cambridge University Press, Cambridge, U. K.

Bibliography 543 Dowsett, R. J. and Dowsett-Lemaire, F. (1980). The systematic status of some Zambian birds. Gerfaut, 70, 151–99.

Laos: two species new for Indochina. Forktail, 11, 159–60.

Dowsett, R. J. and Dowsett-Lemaire, F. (1991). Flore et faune du bassin du Kouilou (Congo) et leur exploitation. Tauraco Research Report, no. 4.

Duckworth, J. W. (1997). Mobbing of a Drongo Cuckoo Surniculus lugubris. Ibis, 139, 190–2.

Dowsett, R. J. and Dowsett-Lemaire, F. (1993). Comments on the taxonomy of some Afrotropical bird species. Tauraco Research Report, no. 5, 323–89.

Duckworth, J. W., Wilkinson, R. J., Tiard, R. J., Kelsh, R. N, Irwin, S. A, Evans, M. I., and Orrell, T. D. (1996). Bird records from Similajau National Park, Sarawak. Forktail, 12, 159–96.

Dowsett, R. J. and Dowsett-Lemaire, F., eds. (1997). Flore et faune du Parc National d’Odzala, Congo. Tauraco Research Report, no. 6.

Dunbar, R. I. M. (1995). Neocortex size and group size in primates: a test of the hypothesis. Journal of Human Evolution, 28, 287–96.

Dowsett, R. J., Aspinwall, D. R., and Leonard, P. M. (1999a). Further additions to the avifauna of Zambia. Bulletin of the British Ornithologists’ Club, 119, 94–103.

Dunn, P. O. and Cockburn, A. (1996). Evolution of male parental care in a bird with almost complete cuckoldry. Evolution, 50, 2542–8.

Dowsett, R. J., Christy, P., and Germain, M. (1999b). Additions and corrections to the avifauna of Central African Republic. Malimbus, 21, 1–15. Dowsett-Lemaire, F. (1990). Ecoethology, distribution and status of Nyungwe Forest birds (Rwanda). Tauraco Research Report, 3, 31–85. Dowsett-Lemaire, F. (1996). Observations of two Cuculus species fed by forest hosts in the Congo. Malimbus, 18, 153–4.

Dunn, P. O., Cockburn, A., and Mulder, R. A. (1995). Fairy-wren helpers often care for young to which they are unrelated. Proceedings of the Royal Society of London, Series B, 259, 339–43. Dunson, W. A., Dunson, M. K., and Ohmart, R. D. (1976). Evidence for the presence of nasal salt glands in the Roadrunner and Coturnix Quail. Journal of Experimental Zoology, 198, 209–16. Durrell, G. (1956). The drunken forest. Rupert HartDavis, London.

Dowsett-Lemaire, F. and Dowsett, R. J. (2000). Birds of the Lobéké Faunal Reserve, Cameroon, and its regional importance for conservation. Bird Conservation International, 10, 67–87.

Durrer, H. (1986). The skin of birds: colouration. In Biology of the integument, vol. 2. Vertebrates, (eds. J. Bereiter-Hahn, A. G. Matoltsky, and K. S. Richards), pp. 239–47. Springer-Verlag, Berlin.

Draffan, R. D. W., Garnett, S. T., and Malone, G. J. (1983). Birds of the Torres Strait: an annotated list and biogeographical analysis. Emu, 83, 207–34.

Durrer, H. and Villiger, W. (1970). Schillerradien des Goldkuckucks (Chrysococcyx cupreus (Shaw)) im Elektronenmikroskop. Zeitschrift für Zellforschung und Mikroskopische Anatomie, 109, 407–13.

Dröscher, L. (1990). A study on radio-tracking of the European Cuckoo (Cuculus canorus canorus). Current Topics in Avian Biology. Proceedings of the International Centennial Meeting of the Deutsche OrnithologenGesellschaft, Bonn 1988, 187–93. Dubs, B. (1992). Birds of southwestern Brazil. BetronaVerlag, Küsnacht, Switzerland. Duckworth, J. W. (1991). Responses of breeding Reed Warblers Acrocephalus scirpaceus to mounts of Sparrowhawk Accipiter nisus, Cuckoo Cuculus canorus and Jay Garrulus glandarius. Ibis, 133, 68–74. Duckworth, J. W. (1996). Moustached Hawk-cuckoo Cuculus vagans and Booted Eagle Hieraeetus pennatus in

Dutson, G. (2001). New distributional ranges for Melanesian birds. Emu, 101, 237–48. Dutson, G. C. L., Evans, T. D., Brooks, T. M., Asane, D. C., Timmins, R. J., and Toledo, A. (1992). Conservation status of birds on Mindoro, Philippines. Bird Conservation International, 2, 303–25. Dybbro, T. (1977). Gogen [The Cuckoo]. Rhodos, Copenhagen. Dyck, J. (1987). Structure and light reflection of green feathers of fruit doves (Ptilinopus spp.) and an imperial pigeon (Ducula coccina). Kongelige Danske Videnskabernes Selskab, Biologisk Skrifter, 30, 1–43.

544 Bibliography Dyer, M., Gartshore, M. E., and Sharland, R. E. (1986). The birds of Nindam Forest Reserve, Kagoro, Nigeria. Malimbus, 8, 2–20.

Eisenberg, J. F. (1981). The mammalian radiations, an analysis of trends in evolution, adaptation, and behavior. University of Chicago Press, Chicago.

Eames, J. C. (1996). Ke Go Nature Reserve – the place of wood. World Birdwatch, 18, 6–8.

Eisenmann, E. (1952). Annotated list of birds of Barro Colorado Island, Panama Canal Zone. Smithsonian Miscellaneous Collections, 117 (5), 1–62.

Eames, J. C., Eve, R., and Tordoff, A. W. (2001). The importance of Vu Quang Nature Reserve, Vietnam, for bird conservation, in the context of the Annamese lowlands endemic bird area. Bird Conservation International, 11, 247–87. Eastman, J. (1991). Black-billed Cuckoo, Yellow-billed Cuckoo. In The atlas of breeding birds of Michigan, (eds. R. Brewer, G. A. McPeek, and R. J. Adams), pp. 232–5. Michigan State University Press, East Lansing.

Eisentraut, M. (1935). Biologische Studien im bolivianischen Chaco. Mitteilungen aus dem Zoologischen Museum in Berlin, 20, 367–443. Eisentraut, M. (1963). Die Wirbeltiere des Kamerungebirges. Verlag Paul Parey, Hamburg and Berlin. Eisentraut, M. (1973). Die Wirbeltierfauna von Fernando Poo und Westkamerun. Bonner Zoologische Monographien 3.

Eates, K. R. (1937). The status of the Koel (Eudynamys scolopaceus L.) in Sind. Journal of the Bombay Natural History Society, 39, 406–14.

Elgood, J. H. (1955). On the status of Centropus epomidis. Ibis, 97, 586.

Eaton, S. W. (1979). Notes on the reproductive behavior of the Yellow-billed Cuckoo. Wilson Bulletin, 91, 154–5.

Elgood, J. H. (1973). Rufous phase Senegal Cuckoos in southwest Nigeria: an illustration of Gloger’s rule. Bulletin of the British Ornithologists’ Club, 93, 173.

Eaton, S. W. and Edwards, E. P. (1947). The Mangrove Cuckoo in interior Tamaulipas, Mexico. Wilson Bulletin, 59, 110–1.

Elgood, J. H., Fry, C. H., and Dowsett, R. J. (1973). African migrants in Nigeria. Ibis, 115, 1–45, 375–411.

Eccles, S. D. (1987). The birds of Sâo Tomé—record of a visit, April 1987 with notes on the rediscovery of Bocage’s Longbill. Malimbus, 10, 207–17. Eck, S. (1976). Die Vögel der Banggai-Inseln, insbesondere Pelengs. Zoologische Abhandlung vom Staatlichen Museum für Tierkunde in Dresden , 34, 53–100. Eck, S. (1977). Ergänzendes über die Vögel der Insel Taliabu, Sula-Insel. Zoologische Abhandlung vom Staatlichen Museum für Tierkunde in Dresden , 34, 127–33. Eck, S. (1978). Über die Vogelbalgpräparation f ür wissenschafftliche Studien. Zoologische Abhandlung vom Staatlichen Museum für Tierkunde in Dresden , 34, vi-xii. Edvardsen, E., Moksnes, A., Røskaft, E., Øien, I. J., and Honza, M. (2001). Egg mimicry in Cuckoos parasitizing four sympatric species of warblers. Condor, 103, 829–37. Edwards, E. A. (1998). Notes on Bearded Scrub Robin nesting. Honeyguide, 44, 91–4. Edwards, E. P. (1957). Yellow-billed Cuckoo nesting in Yucatán. Condor, 59, 69–70. Edwards, S. V. and Naeem, S. (1993). The phylogenetic component of cooperative breeding in perching birds. American Naturalist, 141, 754–89.

Elgood, J. H., et al. (1994). The birds of Nigeria. BOU Check-list No. 4. (2nd edn.) British Ornithologists’ Union, London. Ellegren, H. (2000). Evolution of the avian sex chromosomes and their role in sex determination. Trends in Ecology and Evolution, 15, 188–92. Ellis, D. H., Kepler, C. B., Kepler, A. K., and Teebaki, K. (1990). Occurrence of the Long-tailed Cuckoo Eudynamis taitensis on Caroline Atoll, Kiribati. Emu, 90, 202. Elter, O. (1986). La collezione ornitologica del Museo di Zoologia dell’Università di Torino . Cataloghi VIII. Museo Regionale di Scienze Naturali, Torino. Emlen, J. M. (1974). An urban bird community in Tucson, Arizona: derivation, structure, regulation. Condor, 76, 184–97. Emlen, S. T. (1995). An evolutionary theory of the family. Proceedings of the National Academy of Sciences USA, 92, 8092–9. Emlen, S. T. (1997). Predicting family dynamics in social vertebrates. In Behavioural ecology, 4th edn, (eds. J. R. Krebs and N. B. Davies), pp. 228–53. Blackwell Scientific, Oxford University Press, Oxford.

Bibliography 545 Emlen, S. T. and Wrege, P. H. (1988). The role of kinship in helping decisions among White-fronted Bee-eaters. Behavioral Ecology and Sociobiology, 23, 305–15. Emlen, S. T. and Wrege, P. H. (1989). A test of alternate hypotheses for helping behavior in White-fronted Beeeaters of Kenya. Behavioral Ecology and Sociobiology, 25, 303–19. Emry, R. J. and Thorington, R. W. (1982). Descriptive and comparative osteology of the oldest fossil squirrel, Protosciurus (Rodentia: Sciuridae). Smithsonian Contributions to Paleobiology, 47. Endler, J. A. (1993). The color of light in forests and its implications. Ecological Monographs, 61, 1–27. Engelbach, P. (1932). Les oiseaux du Laos méridional. L’Oiseau et la Revue Française d’Ornithologie, 2, 439–98. Engels, W. L. (1938). Cursorial adaptations in birds. Limb proportions in the skeleton of Geococcyx. Journal of Morphology, 63, 207–17.

Erwee, H. P. (1989). Intense courtship feeding in Dideric Cuckoo. Honeyguide, 35, 116. Étchécopar, R. D. and Hüe, F. (1978). Les oiseaux de Chine, de Mongolie et de Corée, vol. 1, non Passereaux. CNRS, Papeete, Tahiti. Euler, C. (1867). Beiträge zur Naturgeschichte der Vögel Brasiliens. Part 2. Journal für Ornithologie , 15, 217–33. Evans, T. D. (1997). Records of birds from the forests of the East Usambara lowlands, Tanzania, August 1994February 1995. Scopus, 19, 92–108. Evans, T. D. and Timmins, R. J. (1998). Records of birds from Laos during January – July 1994. Forktail, 13, 69–96. Evans, T., Tye, A., Cordeiro, N., and Seddon, N. (1997). Birding in and around the East Usambaras, north-east Tanzania. Bulletin of the African Bird Club, 4, 116–29.

Enomoto, K. (1941). A guide to wild birds, vol. 2. Wild Bird Society of Japan Osaka Chapter, Osaka (in Japanese).

Evans, T. D., Towell, H. C., Timmins, R. J., Thewlis, R. M., Stones, A., Robichaud, W. G., and Barzen, J. (2000). Ornithological records from the lowlands of southern Laos during December 1995 – September 1996, including areas on the Thai and Cambodian borders. Forktail, 16, 29–52.

Eriksen, J. (1991). Birds of the season. Oman Bird News, 10, 11–3.

Ewbank, D. A. (1985). Predation by Senegal Coucal. Honeyguide, 31, 218.

Eriksen, J. (1994). Oman Bird Records Committee decisions. Oman Bird News, 15, 6–7.

Ezra, A. (1927). The nesting of Renauld’s Ground Cuckoo. Avicultural Magazine, 4, 231–2.

Eriksen, J. (1995). Oman Bird Records Committee decisions. Oman Bird News, 16, 93.

Faaborg, J. (1985). Ecological constraints on West Indian bird distributions. Ornithological Monographs, 36, 621–53.

von Erlanger, C. F. (1905). Beiträge zur Vogelfauna Nordostafrikas, III. Journal für Ornithologie , 53, 433–99. Ernst, S. (1992a). Zur Vogelwelt des östlichen Altai. Mitteilungen aus der Zoologischen Museum in Berlin 68, Annalen für Ornithologie , 16, 3–59. Ernst, S. (1992b). Wie ruft der Waldkuckuck (Cuculus saturatus)? Ornithologische Mitteilungen, 44, 16–8.

Faaborg, J. and Winters, J. E. (1979). Winter resident returns and longevity and weights of Puerto Rican birds. Bird-banding, 50, 216–23. Falla, R. A., Sibson, R. B., and Turbott, E. G. (1978). Collins guide to the birds of New Zealand and outlying islands. HarperCollins, Auckland.

Ernst, S. (1996). Zweiter Beigrag zur Vogelwelt des östliches Altai. Mitteilungen aus der Zoologischen Museum in Berlin 68, Annalen für Ornithologie , 20, 123–80.

Fandiño Mariño, J. H. (1986). Análise da comunicação sonora no anu-branco Guira guira (Aves: Cuculidae), avaliações eco-etológicas e evolutivas. Dissertation, Universidade Estadual de Campinas, Campinas, SP (São Paulo), Brazil.

Erritzoe, J. (2000). Working bibliography of cuckoos and turacos of the world. Johannes Erritzoe, Taps, Denmark.

Fandiño Mariño, J. H. (1989). A commicação sonora de anu branco. Avaliações Eco-etológicas e Evolutivas. Editora da UNICA MP, Campinas, SP (São Paulo), Brazil.

Ervin, S. (1989). The nesting of the Dark-billed Cuckoo in the Galápagos. Noticias de Galápagos , 48, 8–10.

Farabaugh, S. M. (1982). The ecological and social significance of duetting. In Acoustic communication in

546 Bibliography birds, vol. 2, (eds. D. E. Kroodsma and E. H. Miller), pp. 85–124. Academic Press, New York. Farquhar, M. R., Lorenz, M., Rayner, J. L., and Craig, A. J. F. K. (1996). Feather ultrastructure and skeletal morphology as taxonomic characters in African sunbirds (Nectariniidae) and sugarbirds (Promeropidae). Journal of African Zoology, 110, 321–31.

Fischer, D. H. (1979). Black-billed Cuckoo (Coccyzus erythropthalmus) breeding in south Texas. Bulletin of the Texas Ornithological Society, 12, 25–6. Fisher, R. A. (1930). The genetical theory of natural selection. Clarendon Press, Oxford.

Feare, C. and Craig, A. (1999). Starlings and mynahs. Princeton University Press, Princeton, New Jersey.

Fisher, C. T. (1981). Specimens of extinct, endangered or rare birds in the Merseyside County Museums, Liverpool. Bulletin of the British Ornithologists’ Club, 101, 276–85.

Feduccia, A. (1996). The origin and evolution of birds. Yale University Press, New Haven.

Fisher, T. and Hicks, N. (2000). A photographic guide to birds of the Philippines. New Holland, London.

Feduccia, A. and Martin, L. D. (1976). The Eocene zygodactyl birds of North America (Aves: Piciformes). Smithsonian Contributions in Paleobiology, 27, 101–10.

Fisk, E. J. (1979). Fall and winter birds near Homestead, Florida. Bird-banding, 50, 224–43.

Fell, H. B. (1947). The migration of the New Zealand Bronze Cuckoo, Chalcites lucidus lucidus (Gmelin). Transactions of the Royal Society of New Zealand, 76, 504–14. Felsenstein, J. ( 1985) Confidence limits on phylogenies: an approach using the bootstrap. Evolution, 39, 783–91. Ferguson, J. W. H. (1994). The importance of low host densities for successful parasitism of Diederik Cuckoos on Red Bishop Birds. South African Journal of Zoology, 29, 70–73. ffrench, R. P. (1991). A guide to the birds of Trinidad and Tobago. (2nd edn.) Livingston, Wynnewood, PA. Fiebig, J. (1993). Dreijährige ornithologische Studien in Nordkorea. 1. Allgemeiner Teil und Non-Passeriformes. Mitteilungen aus der Zoologischen Museum in Berlin, 69, Annalen für Ornithologie 17, 93–146. Fiebig, J. (1995). Dreijährige ornithologische Studien in Nordkorea. Nachtrag zum 1. Allgemeiner Teil (NonPasseres) und 2. Teil Passeriformes. Mitteilungen aus der Zoologischen Museum in Berlin, 71 (Suppl.), Annalen für Ornithologie, 19, 43–99. Finn, P. G. and Hughes, J. M. (2001). Helping behaviour in Australian magpies, Gymnorhina tibicen. Emu, 101, 57–64. Finsch, O. (1898). On the specific distinction of the ground-cuckoos of Borneo and Sumatra (Carpococcyx radiatus and C. viridis). Notes from the Leyden Museum, 20, 97–100. Finsch, O. (1901). Ueber eine dritte sendung Vogelbäge aus Central-Borneo (Mahakkam), gesammelt von Herrn Dr. A. W. Nieuwenhuis. Notes from the Leyden Museum, 22, 163–78.

Fitzgerald, T. D. (1995). The tent caterpillars. Cornell University Press, Ithaca, New York. Fjeldså, J. and Krabbe, N. (1986). Some range extensions and other unusual records of Andean birds. Bulletin of the British Ornithologists’ Club, 106, 115–24. Fjeldså, J. and Krabbe, N. (1990). Birds of the high Andes. Zoological Museum, University of Copenhagen. Flannery, T. F., and Roberts, R. G. (1999). Late Quaternary extinctions in Australasia, an overview. In Extinctions in near time: causes, contexts, and consequences, (ed. R. D. E. MacPhee), pp. 239–55. Kluwer Academic/Plenum Publishers, New York and Dordrecht. Flannery, T. F., and Wickler, S. (1990). Quaternary murids (Rodentia: Mammalia) from Buka Island, Papua New Guinea, with descriptions of two new species. Australian Mammalogy, 13, 127–39. Fleischer, R. C. (2001). Taxonomic and evolutionarily significant unit (ESU) status of western Yellow-billed Cuckoos (Coccyzus americanus). Report to Sacramento Fish and Wildlife Service, Sacramento, California. Fleischer, R. C., Murphy, M. T., and Hunt, L. E. (1985). Clutch size increase and intraspecific brood parasitism in the Yellow-billed Cuckoo. Wilson Bulletin, 97, 125–7. Fleming, R. L. Jr. (1967). The birds of Mussoorie, Uttar Pradesh, India— a distributional and ecological study . PhD Dissertation, Michigan State University, East Lansing. Fleming, R. L. Jr. (1977). Comments on the endemic birds of Sri Lanka. Ceylon Bird Club and Wildlife and Nature Protection Society of Sri Lanka, Colombo.

Bibliography 547 Fleming, R. L. Sr., Fleming, R. L. Jr., and Bangdel, L. S. (1979). Birds of Nepal. (2nd edn.) Avalok, Kathmandu.

Frauca, H. (1974). The natural history of the Pheasant Coucal. Australian Birdlife, 1, 89–93.

Flint, V. E., Boehme, R. L., Kostin, Y. V., and Kuznetsov, A. A. (1984). A field guide to birds of the USSR. Princeton University Press.

Freeman, S. and Zink, R. M. (1995). A phylogenetic study of the blackbirds based on variation in mitochondrial DNA restriction sites. Systematic Biology, 44, 409–20.

Fogden, M. P. L. (1972). The seasonality and population dynamics of equatorial forest birds in Sarawak. Ibis, 114, 307 – 43. Fogden, M. (1976). A census of a bird community in tropical rain forest in Sarawak. Sarawak Museum Journal, 24, 251– 67. Folse, L. J. (1974). Population ecology of Roadrunners (Geococcyx californianus) in south Texas. MS thesis, Texas A and M University, Lubbock. Folse, L. J. and Arnold, K. A. (1976). Secondary sex characteristics in Roadrunners. Bird-banding, 47, 115–8. Folse, L. J. and Arnold, K. A. (1978). Population ecology of Roadrunners (Geococcyx californianus) in south Texas. Southwestern Naturalist, 23, 1–28. Fooden, J. (1969). Taxonomy and evolution of the monkeys of Celebes (Primates: Cercopithecidae). Bibliotheca Primatologica, 10, 1–148. Forbush, E. H. (1927). Birds of Massachusetts, part II. Massachusetts Department of Agriculture, Boston. Ford, J. (1981). Hybridization and migration in Australian populations of the Little and Rufous-breasted Bronze-cuckoos. Emu, 81, 209–22. Ford, J. (1987a). Hybrid zones in Australian birds. Emu, 87, 158–78. Ford, J. (1987b). Taxonomic status of Pheasant Coucal, Red-tailed Black-cockatoo and Red-backed Fairy-wren in the Pilbara, Western Australia. Western Australian Naturalist, 17, 17–8. Foster, D. (1989). Oman Bird Records Committee meetings. Oman Bird News, 7, 3. Foster, M. S. (1975). The overlap of molting and breeding in some tropical birds. Condor, 77, 304 – 14. Franzreb, K. E. and Laymon, S. A. (1993). A reassessment of the taxonomic status of the Yellow-billed Cuckoo. Western Birds, 2, 17–28. Frauca, H. (1967). Birds from the seas, swamps and scrubs of Australia. Heinemann, Melbourne.

Frere, P. J. (1984). White-browed Coucal Centropus superciliosus robbing nest of Grosbeak Weaver Amblyospiza albifrons of eggs. Scopus, 8, 48. Fridolfsson, A.-K., Cheng, H., Copeland, N. G., Jenkins, N. A., Liu, H.-C., Raudsepp, T., Woodage, T., Chowdhary, B., Halverson, J., and Ellegren, H. (1998). Evolution of the avian sex chromosomes from an ancestral pair of autosomes. Proceedings of the National Academy of Sciences, USA, 95, 8147–52. Friedmann, H. (1927). Notes on some Argentine birds. Bulletin of the Museum of Comparative Zoology, Harvard, 68, 139–236. Friedmann, H. (1929). The cowbirds. A study in the biology of social parasitism. Charles C. Thomas, Springfield, Illinois. Friedmann, H. (1930). Birds collected by the Childs Frick Expedition to Ethiopia and Kenya Colony, part 1. Non-passerines. United States National Museum Bulletin, 153. Friedmann, H. (1933). A contribution to the life-history of the Crespín or Four-winged Cuckoo, Tapera naevia. Ibis, Ser. 13, vol. 3, 532–9. Friedmann, H. (1948). The parasitic cuckoos of Africa. Monograph No. 1, Washington Academy of Sciences, Washington. Friedmann, H. (1955). The honey-guides. United States National Museum Bulletin, 208. Friedmann, H. (1956). Further data on African parasitic cuckoos. Proceedings of the United States National Museum, 106, 377–408. Friedmann, H. (1964a). Evolutionary trends in the avian genus Clamator. Smithsonian Miscellaneous Collections, 146 (4), 1–127. Friedmann, H. (1964b). The history of our knowledge of avian brood parasitism. Centaurus, 10, 282–304. Friedmann, H. (1966). A contribution to the ornithology of Uganda. Bulletin of the Los Angeles County Museum of Natural History Science, 3, 1–55.

548 Bibliography Friedmann, H. (1967). Alloxenia in three African species of Cuculus. Proceedings of the United States National Museum, 124, 1–13. Friedmann, H. (1968). The evolutionary history of the avian genus Chrysococcyx. United States National Museum Bulletin, 265. Friedmann, H. (1969). Additions to knowledge of the Yellow-throated Glossy Cuckoo, Chrysococcyx flavigularis. Journal für Ornithologie , 110, 176–80. Friedmann, H. (1978). Result of the Lathrop Central African Expedition 1976, ornithology. Contributions in Science, Natural History Museum of Los Angeles, 287, 1–22. Friedmann, H. and Loveridge, A. (1937). Notes on the ornithology of tropical East Africa. Bulletin of the Museum of Comparative Zoölogy , 81, 1–413. Frith, C. B. (1975). Field observations on Centropus toulou insularis on Aldabra atoll. Ostrich, 46, 251–7. Frome, N. F. (1946). Birds noted in the MahasuNarkanda-Baghi area of the Simla hills. Journal of the Bombay Natural History Society, 46, 308–16. Fry, C. H., Fry, K. and Harris, A. (1992). Kingfishers, beeeaters and rollers. Princeton University Press, Princeton. Fry, C. H., Keith, S., and Urban, E. (eds.). (2000). The birds of Africa, vol. 6. Academic Press, New York. Fuller, R. A. and Erritzøe, J. (1997). Little-known Oriental bird: Red-faced Malkoha Phaenicophaeus pyrrhocephalus. Oriental Bird Club Bulletin, 26, 35–9. Fulton, R. (1903). The Kohoperoa or Koekoe, Long-tailed Cuckoo (Urodynamis taitensis); an account of its habits. Transactions of the New Zealand Institute, 36, 113–48. Fürbinger, M. (1888). Untersuchungen zur Morphologie und Systematik der Vögel, zugleich ein Beitrag zur Anatomie der Stütz- und Bewegungsorgane . Van Holkema, Amsterdam. Gadow, H. (1892). On the classification of birds. Proceedings of the Zoological Society of London, 1892, 229–56.

Solomon Islands. Bulletin of the British Museum (Natural History), Zoology, 9, 1–86. Gallagher, M. and Woodcock, M. W. (1980). The birds of Oman. Quartet Books, London. Gallardo, J. M. (1984). Observaciones sobre el comportamiento del Pirincho (Guira guira) Aves: Cuculiformes. Revista del Museo Argentina de Ciencias Naturales “Bernadino Rivadavia ”, 13, 167–70. Gammie, J. A. (1877). Occasional notes from Sikkim. Stray Feathers, 5, 380–7. García-Moreno, J. and Mindell, D. P. (2000). Rooting a phylogeny with homologous genes on opposite sex chromosomes (gametologs): a case study using avian CHC. Molecular Biology and Evolution, 17, 1826–32. Gargett, V. and Webb, D. G. (1973). A Marico Flycatcher feeds a Didric Cuckoo. Ostrich, 44, 79. Garrett, K. and Dunn, J. (1981). Birds of southern California: status and distribution. Los Angeles Audubon Society, Los Angeles. Garrido, O. H. (1971). Nueva raza del arriero, Saurothera merlini (Aves: Cuculidae) para Cuba. Poeyana, 87, 1–12. Garrido, O. H. and Kirkconnell, A. (2000). Birds of Cuba. Christopher Helm / A. and C. Black, London. Gärtner, K. (1981a). Die Wechselbeziehungen zwischen dem Kuckuck (Cuculus canorus) und dem Sumpfrohrsänger (Acrocephalus palustris) als Beispiel einer Brutparasit-WirtBeziehung. Dissertation, Universität Hamburg. Gärtner, K. (1981b). Das Wegnehmen von Wirtsvogeleiern durch den Kuckuck (Cuculus canorus). Ornithologische Mitteilungen, 33, 115–31. Gärtner, K. (1982). Zur Ablehnung von Eiern und Jungen des Kuckucks (Cuculus canorus) durch die Wirtsvögel—Beobachtungen und experimentelle Untersuchungen am Sumpfrohrsänger (Acrocephalus palustris). Vogelwelt, 103, 201–24. Gartshore, M. E. (1989). An avifaunal survey of Tai National Park, Ivory Coast. BirdLife International, Cambridge, U. K.

Gaines, D. (1974). Review of the status of the Yellowbilled Cuckoo in California: Sacramento Valley populations. Condor, 76, 204–9.

Gaston, A. J. (1976). Brood parasitism by the Pied Crested Cuckoo Clamator jacobinus. Journal of Animal Ecology, 45, 331–48.

Gaines, D. and Laymon, S. A. (1984). Decline, status, and preservation of the Yellow-billed Cuckoo in California. Western Birds, 15, 49–80.

Gaston, A. J. (1977). Social behaviour within groups of Jungle Babblers (Turdoides striatus). Animal Behaviour, 25, 828–48.

Galbraith, I. C. J. and Galbraith, E. H. (1962). Landbirds of Guadalcanal and the San Cristoval group, eastern

Gaston, A. J. (1981). Seasonal breeding, moulting and weight changes among birds of dry deciduous

Bibliography 549 forest in north India. Journal of Zoology, London, 194, 219–43. Gaston, A. J. and Zacharias, V. J. (2000). Hosts of the Common Hawk Cuckoo Hierococcyx varius in India. Forktail, 16, 182. Gatesy, S. M. and Middleton, K. M. (1997). Bipedalism, flight, and the evolution of theropod locomotor diversity. Journal of Vertebrate Paleontology, 17, 308–29. Gatter, W. (1987). Migration behaviour and wintering of Palaearctic birds in Liberia (West Africa). Bird Migration in Liberia I. Verhandlungen der Ornithologischen Gesellschaft in Bayern, 24, 479–508. Gatter, W. (1997). Birds of Liberia. Pica Press, Sussex. Gauthier, J. and Aubry, Y., eds. (1996). The breeding birds of Québec. Province of Quebec Society for the Protection of Birds, Canadian Wildlife Service, Environment Canada, Montréal. Gavrilets, S. and Hastings, A. (1998). Coevolutionary chase in two-species systems with applications to mimicry. Joumnl of Theoretical Biology, 191,415–27. Gehringer, F. (1979). Etude sur le pillage par le coucou, Cuculus canorus, des oeufs de la Rousserolle effarvatte. Nos Oiseaux, 35, 1–16. Geluso, K. N. (1969). Food and survival problems of Oklahoma Roadrunners in winter. Bulletin of the Oklahoma Ornithological Society, 2, 5–6. Geluso, K. N. (1970a). Additional notes on food and fat of Roadrunners in winter. Bulletin of the Oklahoma Ornithological Society, 3, 6. Geluso, K. N. (1970b). Feeding behavior of a Roadrunner in winter. Bulletin of the Oklahoma Ornithological Society, 3, 32. Giai, A. (1949). Sobre un ejemplar joven de Dromococcyx pavoninus Pelzeln. El Hornero, 9, 84 –7. Gibbon, G. (1991). Southern African bird sounds (6 cassettes). Southern African Birding, Hillary.

Gibson-Hill, C. A. (1949b). An annotated checklist of the birds of Malaya. Bulletin of the Raffles Museum, 20, 1–299. Gibson-Hill, C. A. (1949c). Ornithological notes from the Raffles Museum, no. 5. A collection of birds’ eggs from North Borneo. Bulletin of the Raffles Museum, 21, 106–15. Gilbert, B. M., Martin, L. D., and Savage, H. G. (1981). Avian osteology. B. M. Gilbert, Laramie, Wyoming. Gill, B. J. (1980a). Foods of the Shining Cuckoo (Chrysococcyx lucidus, Aves: Cuculidae) in New Zealand. New Zealand Journal of Ecology, 3, 138–40. Gill, B. J. (1980b). Foods of the Long-tailed Cuckoo. Notornis, 27, 96. Gill, B. J. (1982a). Notes on the Shining Cuckoo (Chrysococcyx lucidus) in New Zealand. Notornis, 29, 215–27. Gill, B. J. (1982b). The Grey Warbler’s care of nestlings: a comparison between unparasitized broods and those comprising a Shining Bronze-cuckoo. Emu, 82, 177–81. Gill, B. J. (1983a). Brood-parasitism by the Shining Cuckoo Chrysococcyx lucidus at Kaikoura, New Zealand. Ibis, 125, 40–55. Gill, B. J. (1983b). Morphology and migration of Chrysococcyx lucidus, an Australasian cuckoo. New Zealand Journal of Zoology, 10, 371–82. Gill, E. H. (1923). Plumage display by the Sirkeer Cuckoo (T. leschenaulti). Journal of the Bombay Natural History Society, 29, 299. Gill, F. B. and Stokes, C. (1971). Predation on netted bird by Smooth-billed Ani. Wilson Bulletin, 83, 101–2. Gillard, L. (1987). Southern African bird calls. Gillard Bird Cassettes, Johannesburg.

Gibbs, D. (1990). Indonesia. song cassette, Wildsounds.

Gilliard, E. T. (1950). Notes on a collection of birds from Bataan, Luzon, Philippine Islands. Bulletin of the American Museum of Natural History, 94, 457–504.

Gibbs, H. L., Sorenson, M. D., Marchetti, K., Brooke, M. de L., Davies, N. B., and Nakamura, H. (2000). Genetic evidence for female-specific races of the Common Cuckoo. Nature, 407, 183– 6.

Gilliard, E. T. (1961). Exploring New Britain’s land of fire. National Geographic, 119 (2), 261–92.

Gibson-Hill, C. A. (1949a). Bird and mammal type specimens formerly in the Raffles Museum Collections. Bulletin of the Raffles Museum, 19, 133–98.

Gilliard, E. T. and LeCroy, M. (1961). Birds of the Victor Emanuel and Hindenburg Mountains, New Guinea. Bulletin of the American Museum of Natural History, 123, 1–86.

550 Bibliography Gilliard, E. T. and LeCroy, M. (1966). Birds of the Middle Sepik Region, New Guinea. Results of the American Museum of Natural History Expedition to New Guinea in 1953–1954. Bulletin of the American Museum of Natural History, 132, 247–75.

Goodman, S. M. and Gonzales, P. C. (1990). The birds of Mt. Isarog National Park, southern Luzon, Philippines, with particular reference to altitudinal distribution. Fieldiana Zoology, New Series, 60, Publ. 1415.

Gilliard, E. T. and LeCroy, M. (1967a). Results of the 1958–1959 Gilliard New Britain expedition. 4. Annotated list of birds of the Whiteman Mountains, New Britain. Bulletin of the American Museum of Natural History, 135, 173 –216.

Goodman, S. M. and Jennings, M. C. (1988). Notes on the Saudi Arabian bird collection of I. M. Abdel Magid and S. Halfawi in the Giza Zoological Museum. Sandgrouse, 10, 91– 6.

Gilliard, E. T. and LeCroy, M. (1967b). Annotated list of birds of the Adelbert Mountains, New Guinea. Bulletin of the American Museum of Natural History, 138, 51–82. Giraudoux, P., Degauquier, R., Jones, P. J., Weigel, J., and Isenmann, P. (1988). Avifaune du Niger. Malimbus, 10, 1–141. Giribet, G. and Wheeler, W. C. (1999). On gaps. Molecular Phylogenetics and Evolution, 13, 132–43. Gladstein, D. and Wheeler, W. C. (1996). POY: phylogeny reconstruction via direct optimization. American Museum of Natural History, New York. Glue, D. and Morgan, R. (1972). Cuckoo hosts in British habitats. Bird Study, 19, 187–92. Glutz von Blotzheim, U. N. and Bauer, K. M. (1980). Handbuch der Vögel Mitteleuropas . Band 9. Akademische Verlagsgesellschaft, Wiesbaden.

Goodman, S. M. and Meininger, P. L., eds. (1989). The birds of Egypt. Oxford University Press, Oxford. Goodman, S. M. and Putnam, J. S. (1996). The birds of the eastern slopes of the Réserve Naturelle Intégrale d’Andringitra, Madagascar. Fieldiana Zoology, New Series, 85, Publ. 1480, 171–90. Goodman, S. M. and Ravoavy, F. (1993). Identification of bird subfossils from cave surface deposits at Anjohibe, Madagascar, with a description of a new giant coua (Cuculidae: Couinae). Proceedings of the Biological Society of Washington, 106, 24–33. Goodman, S. M. and Wilmé, L. (2003). Cuculiformes: Coua spp., couas. In The natural history of Madagascar, (eds. S. M. Goodman and J. P. Benstead), pp. 1102–10. University of Chicago Press, Chicago.

Gochfeld, M. (1979). Begging by nestling Shiny Cowbirds: adaptive or maladaptive. Living Bird, 17, 41–50.

Goodman, S. M., Willard, D. E., and Gonzales, P. C. (1995). The birds of Sibuyan Island, Romblon Province, Philippines, with particular reference to elevational distribution and biogeographic affinities. Fieldiana Zoology, New Series no. 82.

Goddard, M. T. and Marchant, S. (1983). The parasitic habits of the Channel-billed Cuckoo Scythrops novaehollandiae in Australia. Australian Birds, 17, 65–72.

Goodman, S. M., Pidgeon, M., Hawkins, A. F. A., and Schulenberg, T. S. (1997). The birds of southeastern Madagascar. Fieldiana Zoology, New Series, 87.

Godfrey, R. (1939). The Black-crested Cuckoo. Ostrich, 10, 21–7.

Goodman, S. M., Hawkins, A. F. A., and Razafimahaimodison, J.-C. (2000). Birds of the Parc National de Majojejy, Madagascar: with reference to elevational distribution. Fieldiana Zoology, New Series no. 97, 175–200.

Gonzales, P. C. (1983). Birds of Catanduanes. Zoological Papers of the Philippine National Museum 2. Good, A. I. (1952). The birds of French Cameroon, Part 1. Sciences Naturelles no. 2, Mémoires de l’Institut Français d’Afrique Noire (Centre du Cameroun). Goodman, S. M. (1993). A reconnaissance of Ile Sainte Marie, Madagascar: the status of the forest, avifauna, lemurs and fruit bats. Biological Conservation, 65, 205–12. Goodman, S. M. and Benstead, J. P., eds. (2003). The natural history of Madagascar. University of Chicago Press, Chicago.

Goodwin, D. (1974). Cuculidae. In Birds of the Harold Hall Expeditions, 1962–1970 , (ed. B. P. Hall), pp. 118–26. London: Trustees of the British Museum (Natural History) Publ. 745. Goodwin, D. (1982). Estrildid finches of the world. British Museum (Natural History)., London. Goodwin, W. (2001). Senegal Coucals and distasteful bush locusts. Honeyguide, 47, 182.

Bibliography 551 Gore, M. E. J. (1990). Birds of The Gambia, an annotated check-list. BOU Check-list No. 3, British Ornithologists’ Union, London.

Graves, G. R. and Zusi, R. L. (1990). Avian body weights from the lower Rio Xingu, Brazil. Bulletin of the British Ornithologists’ Club, 110, 20–5.

Gore, M. E. J. and Gepp, A. R. M. (1978). Las aves del Uruguay. Moscna Hnos., Montevideo.

Graves, G. R., Kratter, A. W., and Bates, J. M. (2002). First nesting record of Black-billed Cuckoo (Coccyzus erythropthalmus) in the Lower Mississippi Valley. Journal of Louisiana Ornithology, 5, 46–8.

Gore, M. E. J. and Won Pyong-Oh. (1971). The birds of Korea. Royal Asiatic Society, Seoul. Gorsuch, D. M. (1932). The Roadrunner. Arizona Wildlife, October 1932, 1–11. Gosler, A. G., Barnett, P. R. and Reynolds, S. J. (2000). Inheritance and variation in eggshell patterning in the Great Tit Parus major. Proceedings of the Royal Society of London B, 267, 2469–73. Gosper, D. (1962). Breeding records of the Koel. Australian Bird Watcher, 17, 226–8.

Graves, J. A. M. and Shetty, S. (2001). Sex from W to Z: evolution of vertebrate sex chromosomes and sex determining genes. Journal of Experimental Zoology, 290, 449–62. Gray, G. R. (1840). A list of the genera of birds, with an indication of the typical species of each genus. R. and J. E. Taylor, London.

Gosper, D. (1964). Observations on the breeding of the Koel. Emu, 64, 39–41.

Gray, J. E. (1846). Catalogue of the specimens and drawings of mammalia and birds of Nepal and Thibet presented by B. H. Hodgson to the British Museum. British Museum, London.

Gosper, D. G. (1997). Aspects of breeding of the Common Koel Eudynamys scolopacea and one of its biological hosts, the Magpie-lark Grallina cyanoleuca. Australian Bird Watcher, 17, 11–19.

Green, A. A. and Carroll, R. W. (1991). The avifauna of Dzahnga-Ndoki National Park and Dzanga-Sangha Rainforest Reserve, Central African Republic. Malimbus, 13, 49–66.

Gosse, P. H. (1847). The birds of Jamaica. John Van Voorst, London. Gould, J. (1845a). (Cuculus optatus.) Proceedings of the Zoological Society of London, 1845, 18–9.

Green, D. J., Cockburn, A., Hall, M. L., Ormond, H., and Dunn, P. O. (1995). Increased opportunities for cuckoldry may be why dominant male fairy-wrens tolerate helpers. Proceedings of the Royal Society of London, Series B, 262, 297–303.

Gould, J. (1845b). Birds of Australia, vol. 4. J. Gould, London.

Greenway, J. C., Jr. (1967). Extinct and vanishing birds of the world. (Revised edn.) Dover, New York.

Gould, J. (1856). Descriptions of two new species of true cuckoos (genus Cuculus as restricted). Proceedings of the Zoological Society of London, 1856, 96.

Greenway, J. C., Jr. (1978). Type specimens of birds in the American Museum of Natural History, part 2. Bulletin of the American Museum of Natural History, 161, 1–305.

Gower, D. J., Kupfer, A., Oommen, O. V., Himstedt, W., Nussbaum, R. A., Loader, S. P., Presswell, B., Müller, H., Krishna, S. B., Boistel, R., and Wilkinson, M. (2002). A molecular phylogeny of ichthyophiid caecilians (Amphibia: Gymnophiona: Ichthyophiidae): out of India or out of South East Asia? Proceedings of the Royal Society of London B, 269, 1563 –70.

Gregory, P. (1995). Further studies of the birds of the Ok Tedi area, Western Province, Papua New Guinea. Muruk, 7, 1–38.

Grantham, M. (2000). Birds of Alas Purwo National Park, East Java. Kukila, 11, 97–121. Granvik, H. (1923). Contributions to the knowledge of the East African ornithology, birds collected by the Swedish Mount Elgon Expedition 1920. Journal für Ornithologie, Sonderheft, 23.

Greig-Smith, P. W. (1976). The composition and habitat preferences of the avifauna of the Mole National Park, Ghana. Bulletin of the Nigerian Ornithologists’ Society, 12, 49–66. Greig-Smith, P. W. (1977). Breeding dates of birds in Mole National Park, Ghana. Bulletin of the Nigerian Ornithologists’ Society, 13, 89–93. Grimes, L. G. (1979). The Yellow-billed Shrike Corvinella corvina: an abnormal host of the Yellow-billed Cuckoo

552 Bibliography Cuculus gularis. Bulletin of the British Ornithologists’ Club, 99, 36–8. Grimes, L. G. (1987). The birds of Ghana. BOU Check-list No. 9, British Ornithologists’ Union, London. Grimmett, R., Inskipp, C., and Inskipp, T. (1999). A guide to the birds of India, Pakistan, Nepal, Bangladesh, Bhutan, Sri Lanka, and the Maldives. Princeton University Press, Princeton. Grinnell, J. (1928). A distributional summation of the ornithology of Lower California. University of California Publications in Zoology, 32. Grinnell, J. and Miller, A. H. (1944). The distribution of the birds of California. Pacific Coast Avifauna 27. Griscom, L. and Greenway, J. C., Jr. (1941). Birds of Lower Amazonia. Bulletin of the Museum of Comparative Zoology, Harvard, 88, 83–344. Grocki, D. R. J. and Johnston, D. W. (1974). Chlorinated hydrocarbon pesticides in North American cuckoos. Auk, 91, 186–8. Groschupf, K. (1987). Status of the Yellow-billed Cuckoo (Coccyzus americanus occidentalis) in Arizona and west Texas. United States Fish and Wildlife Service Report, 20181-86-00731, Phoenix, Arizona.

Gyldenstolpe, N. (1945a). The bird fauna of Rio Juruá in western Brazil. Kungliga Svenska Vetenskapsakademiens Handlingar, Stockholm, 3rd series, 22. Gyldenstolpe, N. (1945b). A contribution to the ornithology of northern Bolivia. Kungliga Svenska Vetenskapsakademiens Handlingar, Stockholm, 3rd series, 23. Gyldenstolpe, N. (1955). Birds collected by Dr. Sten Bergman during his expedition to Dutch New Guinea 1948–1949. Arkiv för Zoologi, Kungliga Svenska Vetenskapsakademien, 3rd series, 8 (2), 183–397 ⫹ 6 plates. Hachisuka, M. (1934). The birds of the Philippine Islands, part III. London, H. F. and G. Witherby. Hachisuka, M. and Udagawa, T. (1950–51). Contributions to the ornithology of Formosa, parts 1, 2. Quarterly Journal of the Taiwan Museum, 3, 187–280; 4, 1–180. Hadden, D. (1981). Birds of the North Solomons. Wau Ecology Institute Handbook No. 8. Wau, Papua New Guinea. Haerlid, A., Janke, A., and Arnason, U. (1997). The mtDNA sequence of the Ostrich and the divergence between paleognathous and neognathous birds. Molecular Biology and Evolution, 14, 754–61. Haffer, J. (1975). Avifauna of northwestern Colombia, South America. Bonner Zoologische Monographien 7.

Grote, H. (1927). Ueber die Verbreitung von Cuculus optatus im europäischen Russland. Ornithologische Monatsberichte, 35, 9–11.

Haffer, J. (1977). A systematic review of the Neotropical ground-cuckoos (Aves, Neomorphus). Bonner Zoologische Beiträge , 28, 48–76.

Groth, J. G. and Barrowclough, G. F. (1999). Basal divergences in birds and the phylogenetic utility of the nuclear RAG-1 gene. Molecular Phylogenetics and Evolution, 12, 115–23.

Haffer, J. (1978). Distribution of Amazon forest birds. Bonner Zoologische Beiträche , 29, 38–78.

Grubh, R. B. and Ali, S. (1974). Birds of Goa. Journal of the Bombay Natural History Society, 73, 42–53. Guangdong Institute of Entomology and Zhongshan University (Kuang-tung sheng k`un ch`ung yen chiu so. Tung wu shih). (1983). Hai-Nan Tao Ti Niao Shou (Birds and Mammals of Hainan Island) (in Chinese). Science Press, Beijing. Gundlach, J. (1874). Neue Beiträge zur Ornithologie Cubas, nach eigenen 30jährigen Beobachtungen zusammelgestellt. Journal für Ornithologie , 22, 113–66. Gurney, J. H. (1897). Cuckoos sucking eggs. Transactions of the Norfolk Society, 6, 568–70.

Haffer, J. (1992). On the “river effect” in some forest birds of southern Amazonia. Boletim do Museu Paraense Emílio Goeldi, Série Zoologia, 8, 217–45. Haffer, J. (1997). Alternative models of vertebrate speciation in Amazonia: an overview. Biodiversity and Conservation, 6, 451–76. Hagemeijer, W. J. M. and Blair, M. J., eds. (1997). The EBCC atlas of European breeding birds: their distribution and abundance. T. and A. D. Poyser, London. Hails, C. (1987). Birds of Singapore. Times Editions, Singapore. Hall, R. (1996). Reconstructing Cenozoic SE Asia. In Tectonic evolution of Southeast Asia, (eds. R. Hall and D. J. Blundell), pp. 153–84. London, Geological Society Special Publication, 106.

Bibliography 553 Hall, R. (1998). The plate tectonics of Cenozoic SE Asia and the distribution of land and sea. In Biogeography and geological evolution of SE Asia, (eds. R. Hall and J. D. Holloway), pp. 99–131. Backhuys, Leiden, The Netherlands.

Harris, M. P. (1973). The Galápagos avifauna. Condor, 75, 265–78.

Hamilton, W. D. (1964). The genetical evolution of social behaviour, II. Journal of Theoretical Biology, 7, 17–52.

Harrison, C. (1978). A field guide to the nests, eggs and nestlings of North American birds. Collins, London.

Hamilton, W. D. (1972). Altruism and related phenomena, mainly in social insects. Annual Review of Ecology and Systematics, 3, 193 –232.

Harrison, C. J. O. (1969). The identification of the eggs of the smaller Indian cuckoos. Journal of the Bombay Natural History Society, 66, 478–88.

Hamilton, W. D. (1990). Mate choice near or far. American Zoologist, 30, 341–52.

Harrison, C. J. O. (1971). Notes on the identification of eggs, egg mimicry and distributional history and the status of the form serratus, in the parasitic Clamator cuckoos. Bulletin of the British Ornithologists’ Club, 91, 126–31.

Hamilton, W. D. (1996). Narrow roads of gene land, vol. 1. Evolution of social behaviour. Oxford University Press, Oxford. Hamilton, W. H. and Hamilton, M. (1965). Breeding characteristics of Yellow-billed Cuckoos in Arizona. Proceedings of the California Academy of Sciences, series 4, 32, 405–32. Hamilton, W. J. and Orians, G. H. (1965). Evolution of brood parasitism in altricial birds. Condor, 67, 361–82. Hamling, H. H. (1937). Notes from Lomagundi District, Southern Rhodesia. Ibis, Ser. 14, vol. 1, 175–7. Hanmer, D. B. (1995). Mensural and moult data on seven species of cuckoo and two species of coucal from Mozambique and Malawi. Honeyguide, 41, 65–102. Hanmer, D. B. (2001). An unhappy Diederik Cuckoo. Honeyguide, 47, 181. Hannecart, F. and Letocart, Y. ( 1980). Oiseaux de Nlle Caledonie et des Loyautes. Vol. I. Les Editions Cardinalis, Nouméa, Nouvelle Calédonie. Hardy, A. and Hardy, J. (1973). Some observations of a Pallid Cuckoo’s early life. South Australian Ornithologist, 26, 87–8. Hardy, J. W., Reynard, G. B., and Coffey, B. B. (1990). Voices of the New World cuckoos and trogons (Cuculidae and Trogonidae). (Rev. edn.) Cassette, Ara Records 11, Gainesville, Florida. Harrington, H. H. (1909). The birds of Burma. Rangoon Gazette, Rangoon. Harris, A. H. and Crews, C. R. (1983). Conklings Roadrunner (Geococcyx conklingi): a subspecies of the California Roadrunner (Geococcyx californianus). Southwestern Naturalist, 28, 407–12.

Harris, M. P. (1982). A field guide to the birds of Galápagos . (2nd edn.) Collins, London.

Harrison, C. J. O. (1973). The zoogeographical dispersal of the genus Chrysococcyx. Emu, 73, 129–33. Harrison, C. J. O. (1982). Cuculiform, piciform and passeriform birds in the Lower Eocene of England. Tertiary Research, 4, 71–81. Harrison, C. J. O. and Walker, C. A. (1977). Birds of the British Lower Eocene. Tertiary Research Special Paper, 3, 1–52. Harrison, J. (1999). A field guide to the birds of Sri Lanka. Oxford University Press, Oxford. Hartert, E. (1891). Katalog der Vögelsammlung im Museum der Senckenbergischen Naturforschenden Gesellschaft. Senckenbergische Naturforschende Gesellschaft, Frankfurt am Main. Hartert, E. (1903). On the birds of the Key and SouthEast Islands, and of Ceram-Laut (part 3). Novitates Zoologicae, 10, 232–54. Hartert, E. (1925a). Review of the genus Cacomantis Müll. Novitates Zoologicae, 32, 164–74. Hartert, E. (1925b). Types of birds in the Tring Museum. B. Types in the general collection. Novitates Zoologicae, 32, 138–57. Hartert, E. (1925c). A collection of birds from New Ireland (Neu Mecklenburg). Novitates Zoologicae, 32, 115–36. Hartert, E. (1926). On the birds of the District of Talasea in New Britain. Novitates Zoologicae, 33, 122–45. Hartert, E. (1930). List of the birds collected by Ernst Mayr. Novitates Zoologicae, 36, 27–8.

554 Bibliography Hartert, E. and Stresemann, E. (1925). Ueber die Indoaustalischen glanzkuckucke (Chalcites). Novitates Zoologicae, 32, 158–63. Hartert, E. and Venturi, S. (1909). Notes sur les oiseaux de la Republique Argentine. Novitates Zoologicae, 6, 156 –299. Hartert, E., Paludan, K., Rothschild, W., and Stresemann, E. (1936). Ornithologische Erbebnisse der Expedition Stein 1931–1932. IV. Die Vögel des Weyland-Gebirges und seines Vorlandes. Mitteilungen aus dem Zoologischen Museum in Berlin, 21, 165–240. Hartlaub, G. (1877). Die Vögel Madagascars . Halle, H. W. Schmidt. Hartman, F. A. (1961). Locomotor mechanisms of birds. Smithsonian Miscellaneous Collections, 143(1), 1–91. Harvey, P. H. and Krebs, J. R. (1990). Comparing brains. Science, 249, 140 –146. Hau, M., Wikelski, M., and Wingfield, J. C. (1998). A Neotropical forest bird can measure the slight changes in tropical photoperiod. Proceedings of the Royal Society of London, Series B, 265, 89–95. Hau, M., Wikelski, M., and Wingfield, J. C. (2000). Visual and nutritional cues stimulate reproduction in a Neotropical bird. Journal of Experimental Zoology, 286, 494 –504.

d’Anjanaharibe-Sud, Madagascar. In: A floral and faunal inventory of the Réserve Spéciale d’Anjanaharibe-Sud, Madagascar: with reference to elevational variation. Fieldiana, Zoology, New Series, 90, 93–127. Hay, W. W., DeConto, R. M., Wold, C. N., Wilson, K. M.,Voigt, S., Schulz, M., Wold, A. R., Dullo, W.-C., Ronov, A. B., Balukhovsky, A. N., and Söding, E. (1999). Alternative global Cretaceous paleogeography. Geological Society of America Special Paper, 332, 1–47. Hayes, F. E. (1995). Status, distribution and biogeography of the birds of Paraguay. American Birding Association Monographs in Field Biology, 1. Hayes, F. E. and Samad, I. (2002). Avifauna of the ‘dragon’s teeth’: the Bocas Islands northern Gulf of Paria, between Venezuela and Trinidad. In Studies in Trinidad and Tobago Ornithology Honouring Richard ffrench, (eds. F. E. Hayes and S. A. Temple), pp. 62–85. Department of Life Sciences, University of the West Indies, St Augustine, Occasional Papers, 11. Hayes, F. E., Scharf, P. A., and Ridgely, R. S. (1994). Austral bird migrants in Paraguay. Condor, 96, 83–97. Heaney, L. R. and Regalado, J. C. (1998). Vanishing treasures of the Philippine rain forest. Field Museum, Chicago. Heather, B. D. and Robertson, H. A. (1997). The field guide to the birds of New Zealand. Oxford University Press, Oxford.

Haverschmidt, F. (1955). Beobachtungen an Tapera naevia und ihren Wirtsvögeln in Surinam. Journal für Ornithologie, 96, 337–43.

Heinrich, G. (1956). Biologische Aufzeichnungen über die Vögel von Halmahera und Batjan. Journal für Ornithologie, 97, 31–40.

Haverschmidt, F. (1961). Der Kuckuck Tapera naevia und seine Wirte in Surinam. Journal für Ornithologie , 102, 353 –9.

Heinrich, G. (1958). Zur Verbreitung und Lebensweise der Vögel von Angola. Systematischer Teil I. Journal für Ornithologie, 99, 322–62.

Haverschmidt, F. and Mees, G. F. (1994). Birds of Suriname. VACO, Paramaribo, Suriname.

Hellebrekers, W. P. J. (1942). Revision of the Penard oölogical collection from Surinam. Zoologische Mededeelingen Ryksmuseum vaan natuurlijke Historie, Leiden, 24, 240–75.

Hawkins, A. F. A. and Goodman, S. M. (1999). Bird community variation with elevation and habitat in parcels 1 and 2 of the Réserve Naturelle Intégrale d’Andohahela, Madagascar. Fieldiana, Zoology, New Series, 94, 175–86. Hawkins, A. F. A. and Goodman, S. M. (2003). Introduction to the birds. In The natural history of Madagascar, (eds. S. M. Goodman and J. P. Benstead), pp. 1019–44. University of Chicago Press, Chicago. Hawkins, A. F. A., Thiollay, J.-M., and Goodman, S. M. (1998). The birds of the Réserve Spéciale

Hellebrekers, W. P. J. and Hoogerwerf, A. (1967). A further contribution to our oological knowledge of the island of Java (Indonesia). Zoologische Verhandelingen, 88, 1–164. Hellmayr, C. E. (1916). Die Avifauna von Timor. In: Haniel, Zoologie von Timor 1 (1), 1–112. Hellmayr, C. E. (1929). A contribution to the ornithology of northeastern Brazil. Field Museum of Natural History, Zoology Series, 255, 235–501.

Bibliography 555 Hendricks, D. P. (1975). Copulatory behavior of a pair of Yellow-billed Cuckoos. Auk, 92, 151.

Hindwood, K. A. (1942). Notes on the Coucal. Emu, 42, 249–52.

Hennig, W. (1966). Phylogenetic systematics. University of Illinois Press, Urbana.

Hindwood, K. A. (1957). The Coucal or Swamppheasant. Australian Museum Magazine, 12, 176–82.

Henry, G. M. (1971). A guide to the birds of Ceylon. (2nd edn.) Oxford University Press, London.

Hindwood, K. A. and McGill, A. R. (1951). The ‘Derra Derra’ 1950 camp-out of the R. A. O. U. Emu, 50, 217–38.

Herbert, E. G. (1924). Nests and eggs of birds in central Siam, part 3. Journal of the Natural History Society of Siam, 6, 293–311. Herreid, C. F. (1960). Roadrunner: a predator of bats. Condor, 62, 67. Herrera, C. M. and Hidalgo, J. (1974). Sobre la presencia invernal de Clamator glandarius en Andalucia. Ardeola, 20, 307–11. Herrick, F. H. (1910). Life and behavior of the cuckoo. Journal of Experimental Zoology, 9, 169–233. Herroelen, P. (1983). To what race belong small wintering Cuckoos (Cuculus canorus) in Zaïre? Revue de Zoologie Africaine, 97, 861–6. Higgins, P. J., ed. (1999). Handbook of Australian, New Zealand and Antarctic birds. vol. 4, Parrots to dollarbirds. Oxford University Press, London. Higuchi, H. (1989). Responses of the Bush Warbler Cettia diphone to artificial eggs of Cuculus cuckoos in Japan. Ibis, 131, 94 –8. Higuchi, H. (1998). Host use and egg color of Japanese cuckoos. In Parasitic birds and their hosts, (eds. S. I. Rothstein and S. K. Robinson), pp. 80 –93. Oxford University Press, Oxford. Higuchi, H. and Payne, R. B. (1986). Nestling and fledgling plumages of Cuculus saturatus horsfieldi and C. poliocephalus poliocephalus in Japan. Japanese Journal of Ornithology, 35, 61–5.

Hinkelmann, C. and Fiebig, J. (2001). An early contribution to the avifauna of Paraná, Brazil. The Arkady Fiedler expedition of 1928/29. Bulletin of the British Ornithologists’ Club, 121, 116–127. Hobbs, J. N. (1990). Interactions of juvenile Pallid Cuckoos with Red-capped Robins and other species. Australian Birds, 23, 66–7. Hockey, P. A. R. (1992). Rare birds in South Africa, 1989–1990. Seventh report of the SAOS Rarities Committee. Birding in Southern Africa, 44, 38–44. Hodgson, W. B. (1839). On Cuculus. Journal of the Asiatic Society of Bengal, 8, 136–7. Hoffmann, A. (1950). Der Indische Cuckuck (Cuculus micropterus Gould). Studien aus Peking und Nanking. Bonner Zoologischer Beitrage, 1, 21–30. Hoffmann, T. W. (1989a). Notes on the status and distribution of some birds in Sri Lanka as listed in S. D. Ripley (1982) “A synopsis of the birds of India and Pakistan, together with those of Nepal, Bhutan, Bangladesh and Sri Lanka.” Journal of the Bombay Natural History Society, 86, 7–16. Hoffmann, T. W. (1989b). The Greenbilled or Ceylon Coucal Centropus chlororhynchus Blyth—Sri Lanka’s rarest endemic species. Journal of the Bombay Natural History Society, 86, 339–43. Hoffman, T. W. (1996). New bird records in Sri Lanka and some connected matters. Journal of the Bombay Natural History Society, 93, 382–7.

Higuchi, H. and Sato, S. (1984). An example of character release in host selection and egg colour of cuckoos Cuculus sp. in Japan. Ibis, 126, 398–404.

Holman, F. C. (1947). Birds of the Gold Coast. Ibis, 89, 623–50.

Hildebrand, M. (1974). Analysis of vertebrate structure. John Wiley and Sons, New York.

Holmes, D. A. (1995). Review of MacKinnon and Phillipps 1993. Kukila, 7, 158–64.

Hillis, D. M., Moritz, C., and Mable, B. (eds.). (1996). Molecular systematics, 2nd edn. Sinauer, Sunderland, Massachusetts.

Holmes, D. A. (1997). Kalimantan bird report-2. Kukila, 9, 141–69.

Hilty, S. L. and Brown, W. L. (1986). A guide to the birds of Colombia. Princeton University Press, Princeton.

Holmes, D. A. and Burton, K. (1987). Recent notes on the avifauna of Kalimantan. Kukila, 3, 2–32.

Holmes, D. A. (2000). Editorial. Kukila, 11, 1–2.

556 Bibliography Holmes, D. A. and Phillipps, K. (1996). The birds of Sulawesi. Oxford University Press, Oxford. Holmes, D. and van Balen, S. (1996). The birds of Tinjil and Deli islands, West Java. Kukila, 8, 117–30.

Sandcoleiformes new order). In Papers in Avian Paleontology Honoring Pierce Brodkorb, (ed. K. E. Campbell), Natural History Museum of Los Angeles County Science Series, 36, 137–60.

Holyoak, D. T. (1980). Guide to Cook Island birds. D. T. Holyoak, Rarotonga.

Howard, H. (1929). The avifauna of Emeryville shellmound. University of California Publications in Zoology, 32, 301–94.

Honza, M., Picman, J., Grim, T., Novák, T., Novák, V., Capek, M., and Mrlik, V. (2001). How to hatch from an egg of great structural strength. A study of the Common Cuckoo. Journal of Avian Biology, 32, 249–55.

Howard, H. (1931). A new species of road-runner from Quaternary cave deposits in New Mexico. Condor, 33, 206–9.

Hoogerwerf, A. (1948). Contribution to the knowledge of the distribution of birds on the island of Java, with remarks on some new birds. Treubia, 19, 83–137. Hoogerwerf, A. (1949). Bijdrage tot de Oölogie van Java. Limosa, 22, 1–279. Hoogerwerf, A. (1964). On the species of Centropus from the Kangean Islands (Aves, Cuculidae). Zoologische Mededeelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 40 (14), 119–23. Hoogerwerf, A. (1969). Opmerkingen over het voedsel van Caloenas nicobarica et Ducula bicolor. Ardea, 57, 86–8. Hopkinson, E. (1922). Cuckoos in the Gambia. Ibis, Ser. 18, vol. 4, 744–5. Hoppe, D. (1988). Zucht des Fasankuckucks. Gefiederte Welt, 112, 113–4. Hopwood, J. C. and Mackenzie, J. M. D. (1917). A list of birds from the North Chin Hills. Journal of the Bombay Natural History Society, 25, 72–91. Hornbuckle, J. (1997). Two sightings of Banded Ground-cuckoo Neomorphus radiolosus in Ecuador. Cotinga, 8, 90. Hornskov, J. (1995). Recent observations of birds in the Philippine Archipelago. Forktail, 11, 1–10. Hosono, T. (1969). A study of the life history of Blue Magpie 6. Distribution and movements in Nagano area. Miscellaneous Reports of the Yamashina Institute for Ornithology, 5, 659–75. Houde, P. and Olson, S. L. (1988). Small arboreal nonpasserine birds from the Early Tertiary of western North America. Acta XIX Congressus Internationalis Ornithologici, vol. 2, Ottawa, 2030–6. Houde, P. and Olson, S. L. (1992). A radiation of colylike birds from the Eocene of North America (Aves:

Howell, A. H. (1932). Florida bird life. Florida Department of Game and Fresh Water Fish, United States Department of Agriculture, New York. Howell, S. N. G. and Webb, S. (1995). A guide to the birds of Mexico and Northern Central America. Oxford University Press, Oxford. Howell, T. R. (1957). Birds of second-growth rain forest area of Nicaragua. Condor, 59, 73–111. Hubbard, J. P. and Crossin, R. S. (1974). Notes on northern Mexican birds, an expedition report. Nemouria, 14, 1–41. Hudson, W. H. (1874). Notes on the procreant instincts of the three species of Molothrus found in Buenos Ayres. Proceedings of the Zoological Society of London, 1874, 153–74. Hudson, W. H. (1920). Birds of La Plata. 2 vols. Dent and Sons, London. Huels, T. R. (1982). Co-operative feeding of conspecific and Clamator jacobinus young by Turdoides rubiginosus. Scopus, 6, 33–5. Hughes, J. M. (1996a). Phylogenetic analysis of the Cuculidae (Aves, Cuculiformes) using behavioral and ecological characters. Auk, 113, 10–22. Hughes, J. M. (1996b). Greater Roadrunner Geococcyx californianus. In The Birds of North America No. 244, (eds. A. Poole and F. Gill). The Academy of Natural Sciences, Philadelphia, PA, and the American Ornithologists’ Union, Washington, D.C. Hughes, J. M. (1997a). Taxonomic significance of hostegg mimicry by facultative brood parasites of the avian genus Coccyzus (Cuculidae). Canadian Journal of Zoology, 75, 1380–6. Hughes, J. M. (1997b). Mangrove Cuckoo Coccyzus minor. In The Birds of North America no. 299, (eds. A. Poole and F. Gill). The Academy of Natural Sciences,

Bibliography 557 Philadelphia, and the American Ornithologists’ Union, Washington, D.C.

Körpergewichtes beim Kuckuck Cuculus canorus. Ö kologie der Vogel , 2, 130–1.

Hughes, J. M. (1997c). Vocal duetting by a mated pair of Coral-billed Ground-cuckoos (Carpococcyx renauldi ) at the Metro Toronto Zoo. Zoo Biology, 16, 179–86.

Hunter, H. C. (1961). Parasitism of the Masked Weaver Ploceus velatus arundinaceus. Ostrich, 32, 55–63.

Hughes, J. M. (1997d). Systematics of New World Cuckoos (Aves, Cuculidae) and the evolution of brood parasitism. PhD Dissertation, University of Toronto. Hughes, J. M. (1999a). Yellow-billed Cuckoo Coccyzus americanus. In The Birds of North America no. 418, (eds. A. Poole and F. Gill). The Academy of Natural Sciences, Philadelphia, and the American Ornithologists’ Union, Washington, D.C.

Hustler, K. (1985). Barred Long-tailed Cuckoo in Mana Pools National Park. Honeyguide, 31, 171. Hustler, K. (1997a). Notes on the breeding biology of the Coppery-tailed Coucal. Honeyguide, 44, 184–7. Hustler, K. (1997b). The status, breeding and parasitism of the White-rumped Babbler in Zimbabwe. Honeyguide, 43, 211–3.

Hughes, J. M. (1999b). A closer look: Mangrove Cuckoo. Birding, 31 (1), 22–7.

Hustler, K., Best, G., and Voss, N. (1996). First breeding record of the Coppery-tailed Coucal in Zimbabwe. Honeyguide, 42, 163–5.

Hughes, J. M. (2000). Monophyly and phylogeny of cuckoos (Aves, Cuculidae) inferred from osteological characters. Zoological Journal of the Linnean Society, 130, 263–307.

Hutchinson, J. N. (1983). Australian bird calls. 2nd edn., cassette. John N. Hutchinson, Balingup, Western Australia.

Hughes, J. M. (2001). Black-billed Cuckoo Coccyzus erythropthalmus. In The Birds of North America no. 587, (eds. A. Poole and F. Gill). The Academy of Natural Sciences, Philadelphia, and the American Ornithologists’ Union, Washington, D.C. Hughes, J. M. and Baker, A. J. (1999). Phylogenetic relationships of the enigmatic Hoatzin (Opisthocomus hoazin) resolved using mitochondrial and nuclear gene sequences. Molecular Biology and Evolution, 16, 1300–7. Hume, A. O. (1873). Nest and eggs of Indian birds, rough draft. Office of Superintendent of Government Printing, Calcutta. Hume, A. O. (1874). Contributions to the ornithology of India, the islands of the Bay of Bengal. Stray Feathers, 2, 29–324. Hume, A. O. (1875). A first list of the birds of Upper Pegu. Stray Feathers, 3, 1–194. Hume, A. O. (1888). The birds of Manipur, Assam, Sylhet and Cachar. Stray Feathers, 11. Hume, A. O. and Davison, W. (1878). A revised list of the birds of Tenasserim. Stray Feathers, 6. Hume, A. O. and Oates, E. W. (1890). The nests and eggs of Indian birds, vol. 2. (2nd edn.) R. H. Porter, London. Hund, K. and Prinzinger, R. (1980). Zur Jugendentwicklung der Körpertemperature und des

Huxley, T. H. (1867). On the classification of birds; and on the taxonomic value of the modifications of certain of the cranial bones observable in that class. Proceedings of the Zoological Society of London, 1867, 415–72. Huxley, T. H. (1868). On the classification and distribution of the Alectoromorphae and Heteromorphae. Proceedings of the Zoological Society of London, 1868, 294–319. Hwang, Bo-Yeon. (1997). Songs of Korean birds. Laboratory of Ornithology, Kyung-Hee University, Seoul, Korea. Iapichino, C. and Massa, B. (1989). The birds of Sicily. BOU Check-list No. 11, British Ornithologists’ Union, London. ICZN (International Commission on Zoological Nomenclature). (1985). International code of zoological nomenclature, 3rd Edition. International Trust for Zoological Nomenclature, London. ICZN (International Commission on Zoological Nomenclature). (1999). International code of zoological nomenclature, 4th Edition. International Trust for Zoological Nomenclature, London. Immelmann, K. and Immelmann, G. (1967). Verhaltensökologische Studien an afrikanischen und australischen Estrildiden. Zoologische Jahrbücher. Abteilung für Systematik, Geographe und Biologie der Tiere , 94, 609–86.

558 Bibliography Immelmann, K. (1971). Ecological aspects of periodic reproduction. In Avian biology, vol. 1, (eds. D. S. Farner and J. R. King), pp. 341–89. Academic Press, New York. Inglis, C. M. (1903). The birds of the Madhubani subdivision of the Darbhanga District, Tirhut, with notes on species noticed elsewhere in the District. Part 3. Journal of the Bombay Natural History Society, 14, 362–71. Inglis, C. M. (1908). The oology of Indian parasitic cuckoos. Journal of the Bombay Natural History Society, 18, 681–2. Inglis, C. M. (1923). The fauna of the Darbhanga District, North Bihar. Journal of the Bombay Natural History Society, 29, 136–45. Inglis, C. M., Travers, W. L., O’Donel, H. V., and Shebbeare, E. O. (1920). A tentative list of the vertebrates of the Jalpaighuri district, Bengal. Part 2. Birds. Journal of the Bombay Natural History Society, 26, 988–99. Inskipp, C. and Inskipp, T. (1985). A guide to the birds of Nepal. Smithsonian Institution Press, Washington, D.C. Inskipp, C., Inskipp, T., and Sherub. (2000). The ornithological importance of Thrumshingla National Park, Bhutan. Forktail, 16, 147–62. Inskipp, T., Lindsey, N., and Duckworth, W. (1996). An annotated checklist of the birds of the Oriental Region. Oriental Bird Club, Sandy, U.K. Irisov, E. I. (1967). [The Himalayan Cuckoo in the south-eastern Altai]. Ornitologiya, 8, 355–6. Irwin, D. E. and Price, T. (1999). Sexual imprinting, learning and speciation. Heredity, 82, 347–54. Irwin, M. P. S. (1956). Field notes on a collection from Mozambique. Ostrich, 27, 28–39. Irwin, M. P. S. (1981). The birds of Zimbabwe. Quest, Harare. Irwin, M. P. S. (1982). A fourth specimen of the Lesser Cuckoo from Zimbabwe. Honeyguide, 111/112, 63–4. Irwin, M. P. S. (1985). Interrelationships among African species of Centropus (Cuculidae). Ostrich, 56, 132–4. Irwin, M. P. S. (1987a). Earliest record of the Barred Long-tailed Cuckoo from the Middle Zambezi Valley. Honeyguide, 33, 150. Irwin, M. P. S. (1987b). Geographical variation in plumage of female Klaas Cuckoo. Malimbus, 9, 43–6.

Irwin, M. P. S. (1988). Cuculidae, cuckoos, malkohas and coucals. In The birds of Africa, vol. 3, (eds. C. H. Fry, S. Keith and E. K. Urban), pp. 58–104. Academic Press, London. Irwin, M. P. S. (1989). Vagrant cuckoos in the southern Indian Ocean: a comment. Cormorant, 17, 87–8. Irwin, M. P. S. and Benson, C. W. (1966). Notes on the birds of Zambia. Arnoldia (Rhodesia), 32 (2), 1–19. Isenmann, P. and Moali, A. (2000). Oiseaux d’Algérie/ Birds of Algeria. SEOF/Société d’Études Ornithologiques de France, Paris. Iwaniuk, A. N. and Nelson, J. E. (2002). Can endocranial volume be used as an estimate of brain size in birds? Canadian Journal of Zoology, 80, 16–23. Iwaniuk, A. N. and Arnold, K. E. (2004). Is cooperative breeding associated with bigger brains? A comparative test in the Corvida (Passeriformes). Ethology, 110, 203–20. Jackson, F. J. (1938). The birds of Kenya Colony and the Uganda Protectorate, vol. 1. Gurney and Jackson, London. Jackson, W. M. (1998). Egg discrimination and egg-color variability in the Northern Masked Weaver, the importance of conspecific versus interspecific parasitism. In Parasitic birds and their hosts, (eds. S. I. Rothstein and S. K. Robinson), pp. 407–16. Oxford University Press, Oxford. Jacobs, H. and Martron, N. (1996). Asian-Pacific birds: Pied Cuckoo on Phuket Island, Thailand, in January 1995. Dutch Birding, 18, 22. Jaeger, E. C. (1947). Stone-turning habits of some desert birds. Condor, 49, 171. Jahn, A. E., Davis, S. E., and Zankys, A. M. S. (2002). Patterns of austral bird migration in the Bolivian chaco. Journal of Field Ornithology, 73, 258–67. Jamdar, N. (1987). Occurrence of Pied Crested Cuckoo (Clamator jacobinus) in Suru Valley, Ladakh. Journal of the Bombay Natural History Society, 84, 208–9. James, D. A. and Neal, J. C. (1986). Arkansas birds: their distribution and abundance. University of Arkansas Press, Fayetteville. James, H. F. and Olson, S. L. (1991). Descriptions of thirtytwo new species of birds from the Hawaiian Islands: Part II. Passeriformes. Ornithological Monographs, 46.

Bibliography 559 James, H. W. (1970). Catalogue of the birds eggs in the collection of the National Museums of Rhodesia. Queen Victoria Museum Special Publication, Salisbury.

Jerdon, T. C. (1862). The birds of India being a natural history of all the birds known to inhabit continental India, vol. 1. The author, Military Orphan Press, Calcutta.

Jany, E. (1955). Neue Vogel-Formen von den NordMolukken. Journal für Ornithologie , 96, 102–6.

Jerdon, T. C. (1873). Supplementary notes to ‘The Birds of India.’ Ibis, 3rd Ser., vol. 5, 1–22.

Janisch, M. (1954). Fight between Cuculus c. canorus L.—Cuckoo—and Acrocephalus a. arundinaceus L.—Great Reed Warbler. Aquila, 55–58 , 291.

Jerrison, J. H. (1973). Evolution of the brain and intelligence. Academic Press, New York.

Jansen, A. (1990). Large-billed Scrubwrens cooperatively rear a cuckoo. Australian Bird Watcher, 13, 198–9. Janzen, D. H. (1974). Tropical blackwater rivers, animals and mast fruiting by the Dipterocarpaceae. Biotropica, 6, 69–103. Jauvin, D. (1996). Yellow-billed Cuckoo. In The breeding birds of Québec, (eds. J. Gauthier and Y. Aubry), pp. 582–3. Province of Quebec Society for the Protection of Birds, Canadian Wildlife Service, Environment Canada, Montréal. Javed, S. and Rahmani, A. R. (1998). Conservation of the avifauna of Dudwa National Park, India. Forktail, 14, 55–64. Jenner, E. (1788). Observations on the natural history of the Cuckoo. Philosophical Transactions of the Royal Society of London, 78, 219–35. Jennings, M. C. (1981). The birds of Saudi Arabia: a checklist. Cambridge, U. K. Jenny, J. (1997). Unusual breeding behaviour of the Guira Cuckoo Guira guira. Cotinga, 7, 18. Jensen, R. A. C. (1975). A tale of two cuckoos. Bokmakierie, 27, 29–30. Jensen, R. A. C. (1980). Cuckoo egg identification by chromosome analysis. Proceedings of the VI Pan-African Ornithological Congress, 23–5. Jensen, R. A. C. and Clinning, C. F. (1974). Breeding biology of two cuckoos and their hosts in South West Africa. Living Bird, 13, 5–50. Jensen, R. A. C. and Jensen, M. K. (1969). On the breeding biology of southern African cuckoos. Ostrich, 40, 163–81. Jensen, R. A. C. and Vernon, C. J. (1970). On the biology of the Didric Cuckoo in southern Africa. Ostrich, 41, 237–46.

Jobling, J. A. (1991). A dictionary of scientific bird names. Oxford University Press, Oxford. Johansen, H. (1952). Ornithology in Russia. Ibis, 94, 1–48. Johansen, H. (1955). Die Vogelfauna Westsibiriens. III. Systematik und Verbreitung, Oekologie und Lebensweise der Non-Passeres, Pici—Cuculi. Journal für Ornithologie, 96, 382–410. Johns, A. D. (1989). Recovery of a peninsular Malaysian rainforest avifauna following selective timber logging: the first twelve years. Forktail, 4, 89–105. Johnsingh, A. J. T. and Paramanandham, K. (1982). Group care of White-headed Babblers Turdoides affinis for a Pied Crested Cuckoo Clamator jacobinus chick. Ibis, 124, 179–83. Johnsgard, P. A. (1997). The avian brood parasites. Oxford University Press, Oxford. Johnson, A. W. (1967). The birds of Chile and adjacent regions of Argentina, Bolivia and Peru. A. W. Johnson, Buenos Aires. Johnson, A. W. (1972). Supplement to The birds of Chile and adjacent regions of Argentina, Bolivia and Peru. A. W. Johnson, Buenos Aires. Johnson, C. N. (1983). Courtship feeding and mating in the Channel-billed Cuckoo. Corella, 7, 44. Johnson, K. P. and Lanyon, S. M. (1999). Molecular systematics of the grackles and allies, and the effect of additional sequence (cyt b and ND2). Auk, 116, 759–68. Johnson, K. P., Goodman, S. M., and Lanyon, S. M. (2000). A phylogenetic study of the Malagasy couas with insights into cuckoo relationships. Molecular Phylogenetics and Evolution, 14, 436–44. Johnson, P. G. (1968). Notes on nesting and diet of Burchell’s Coucal. Ostrich, 39, 194.

560 Bibliography Johnston, R. F. (1964). The breeding birds of Kansas. University of Kansas Publications Museum of Natural History, 12, 575–655. Johnstone, R. A., Woodroffe, R., Cant, M. A., and Wright, J. (1999). Reproductive skew in multimember groups. American Naturalist, 153, 315–31. Johnstone, R. E. and Burbidge, A. H. (1991). The avifauna of Kimberley rainforests. In Kimberley rainforests, (eds. N. L. McKenzie, R. B. Johnston and P. G. Kendrick), pp. 361—91. Surrey Beatty, Sydney. Johnstone, R. E. and Storr, G. M. (1998). Handbook of Western Australian birds, vol. 1. Western Australian Museum, Perth. Johnstone, R. E., van Balen, S., and Dekker, R. W. R. J. (1993). New bird records for the island of Lombok. Kukila, 6, 124–7. Jones, A. E. (1941). Presumptive evidence of the nidification of the Indian Cuckoo (Cuculus micropterus Gould). Journal of the Bombay Natural History Society, 42, 931–3. Jones, D. A., Gibbs, H. L., Matsuda, T., Brooke, M. de L., Uchida, H., and Bayliss, M. J. (1997). The use of DNA fingerprinting to determine the possible mating system of an obligate brood parasitic bird, the cuckoo Cuculus canorus. Ibis, 139, 560–2. Jones, D. N., Dekker, R. W. R. J., and Roselaar, C. S. (1995). The megapodes. Oxford University Press, Oxford. Jones, J. M. B. (1985). Striped Crested Cuckoo parasitizing Arrow-marked Babbler. Honeyguide, 31, 170–1. Jones, J. M. B. (1987). Another new locality record for the Barred Long-tailed Cuckoo. Honeyguide, 33, 61–2. Jones, J. M. B. (1992). Striped Cuckoo and Arrowmarked Babbler observations. Honeyguide, 38, 75–6. Jones, J. M. B. (1996). Thick-billed Cuckoo in Marondera North. Honeyguide, 42, 31.

Jubb, R. A. (1952). Some notes on birds of Southern Rhodesia. Ostrich, 23, 162–4. Jullien, M. and Thiollay, J.-M. (1998). Multi-species territoriality and dynamic of Neotropical forest understorey bird flocks. Journal of Animal Ecology, 67, 227–52. Junge, G. C. A. (1936). Fauna Simalurensis-Aves. Temminckia, 1, 1–76. Junge, G. C. A. (1937a). Further notes on the birds of Simalur. Temminckia, 2, 198–202. Junge, G. C. A. (1937b). On Bonaparte’s types of the cuckoos belonging to the genus Piaya. Zoologische Mededeelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 19, 183–5. Junge, G. C. A. (1937c). The birds of south New Guinea. Part 1. Non Passeres. Nova Guinea, New Series 1, 125–88. Junge, G. C. A. (1938). Remarks on Chalcites malayanus (Raffles). Zoologische Mededeelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 20, 237–9. Junge, G. C. A. (1948). Notes on some Sumatran birds. Zoologische Mededeelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 29, 311–26. Junge, G. C. A. (1953). Zoological results of the Dutch New Guinea Expedition, 1939, no. 5. Zoologische Verhandelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 20, 1–77. Junge, G. C. A. (1956). On Cuculus canoroïdes S. Müller. Auk, 73, 555–6. Junge, G. C. A. and Kooiman, J. G. (1951). On a collection of birds from the Khwae Noi valley, western Siam. Zoologische Verhandelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 15, 1–38. Kaempfer, E. (1924). Ueber das Vogelleben in Santo Domingo. Journal für Ornithologie , 72, 178–84. Kavanaugh, J. L. and Ramos, J. (1970). Roadrunners: activity of captive individuals. Science, 169, 780–2.

Joseph, L., Wilke, T., and Alpers, D. (2002). Reconciling genetic expectations from host specificity with historical population dynamics in an avian brood parasite, Horsfield’s Bronze-cuckoo Chalcites basalis of Australia. Molecular Ecology, 11, 829–37.

Keast, A., Recher, H. F., Ford, H. A., and Saunders, D., eds. (1985). Birds of eucalypt forests and woodlands: ecology, conservation, management. RAOU, Surrey Beatty, Sydney.

Jourdain, F. C. R. (1925). A study of parasitism in the cuckoos. Proceedings of the Zoological Society of London, 1925, 639–67.

Keith, G. S. (1968). Notes on birds of East Africa, including additions to the avifauna. American Museum Novitates, 2321.

Bibliography 561 Keith, G. S. and Gunn, W. W. H. (1971). Birds of the African rain forest. Sounds of Nature, no. 9. Two sound disks.

Kilner, R. M. and Davies, N. B. (1999). How selfish is a cuckoo chick? Animal Behaviour, 58, 797–808.

Keith, G. S. and Vernon, C. J. (1969). Bird notes from northern and eastern Zambia. Puku, 5, 131–9.

King, B. (2002). The Hierococcyx fugax, Hodgson’s Hawk Cuckoo, complex. Bulletin of the British Ornithologists’ Club, 122, 74–80.

Kelly, C. (1987). A model to explore the rate of spread of mimicry and rejection in hypothetical populations of cuckoos and their hosts. Journal of Theoretical Biology, 125, 283–99. Kelsall, H. J. (1914). Notes on a collection of birds from Sierra Leone. Ibis, Series 10, vol. 2, 192–228. Kemp, A. C., Kemp, M. I., Jensen, R. A. C., and Clinning, C. F. (1972). Records of brood parasitism from central South West Africa. Ostrich, 43, 145–8. Kemp, N. (2000). The birds of Siberut, Mentawai Islands, West Sumatra. Kukila, 11, 73–96. Kennedy, R. S., Gonzales, P. C., Dickinson, E. C., Miranda, H. C., and Fisher, T. H. (2000). A guide to the birds of the Philippines. Oxford University Press, Oxford. Kennerley, P. R. and Leader, P. J. (1991). Separation of Cuckoo and Oriental Cuckoo. Dutch Birding, 13, 143–5. Kent, T. H. (1988). Mapping vagrants. Iowa Bird Life, 58, 101–5. Kepler, C. B. and Kepler, A. K. (1978). Status and nesting of the Yellow-billed Cuckoo in Puerto Rico. Auk, 95, 417–9. Khajuria, H. (1984). The Crow-pheasant, Centropus sinensis (Stephens) (Aves: Cuculidae) of central and eastern Madhya Pradesh. Records of the Zoological Survey of India, 81, 89–93. Khayutin, S. N., Dmitrieva, L. P., Tartygina, N. G., and Alesandrov, L. I. (1982). [The behaviour of a nestling Cuculus canorus in the nest of Phoenicurus phoenicurus]. Zoologischeskii Zhurnal, 61, 1063–77. Kiff, L. F. and Williams, A. (1978). Host records for the Striped Cuckoo from Costa Rica. Wilson Bulletin, 90, 138–9. Kikkawa, J. (1968). Courtship feeding in cuckoos. Emu, 8, 213–4. Kikkawa, J. and Dwyer, P. D. (1962). Who feeds the fledged Pallid Cuckoo? Emu, 62, 169–71. Kilgore, D. L., Bernstein, M. H., and Hudson, D. M. (1976). Brain temperatures in birds. Journal of Comparative Physiology, 110, 209–15.

King, B. F. and Dickinson, E. C. (1975). A field guide to the birds of South-east Asia. Houghton Mifflin, Boston. King, B., Buck, H., Ferguson, R., Fisher, T., Goblet, C., Nickel, H., and Suter, W. (2001). Birds recorded during two expeditions to north Myanmar (Burma). Forktail, 17, 29–40. King, B. (2005). The taxonomic status of three subspecies of Cuculus saturatus. Bulletin of the British Ornithologists’ Club, 125, in press. Kinnaird, M. F., O’Brien, T. G., and Suryadi, S. (1996). Population fluctuation in Sulawesi Red-knobbed Hornbills: tracking figs in space and time. Auk, 113, 431–40. Kinnaird, M. F., O’Brien, T. G., and Suryadi, S. (1999). The importance of figs to Sulawesi’s imperiled wildlife. Tropical Biodiversity, 6, 5–18. Kinnear, N. B. (1934). On the birds of the Adung Valley, north-east Burma. Journal of the Bombay Natural History Society, 37, 347–68. Kirwan, G. M. and Sharpe, C. J. (1999). Range extensions and notes on the status of little-known species from Venezuela. Bulletin of the British Ornithologists’ Club, 119, 38–47. Kirwan, G. M., Martins, R. P., Morton, K. M., and Showler, D. A. (1996). The status of birds in Socotra and ‘Abd Al-Kuri and the records of the OSME survey in spring 1993. Sandgrouse, 17, 83–101. Kiyosu, Y. (1952). The birds of Japan, II. Kodansha, Tokyo. Kizungu, B. (2000). Observations d’un nid du coucal noire Centropus monachus et attitude de l’homme face à ses oisillons à Irangi, République Démocratique du Congo. Malimbus, 22, 88–90. Klaas, E. E. (1968). Summer birds from the Yucatan peninsula, Mexico. University of Kansas Publications Museum of Natural History, 17, 579–611. Klapste, J. (1981). Observation of one Pallid Cuckoo feeding another. Australian Bird-Watcher, 9, 27–8.

562 Bibliography Klein, J. and Takahata, N. (2002). Where do we come from? The molecular evidence for human descent. Springer Verlag, Berlin. Klein, N. K. and Payne, R. B. (1998). Evolutionary associations of brood parasitic finches (Vidua) and their host species: analyses of mitochondrial restriction sites. Evolution, 52, 299–315. Klicka, J. and Zink, R. M. (1997). The importance of Recent ice ages in speciation: a failed paradigm. Science, 277, 1666–9. Klicka, J., Johnson, K. P. and Lanyon, S. M. (2000). New-World nine-primaried oscine relationships: constructing a mitochondrial DNA framework. Auk, 117, 321–36. Kloot, T. (1993). Photographing the Channel-billed Cuckoo Scythrops novaehollandiae. Bird Observer, 735, 7. Kloss, C. B. (1918). On birds recently collected in Siam. Part I. Phasianidae—Eurylaemidae. Ibis, Ser. 10, vol. 6, 76–114. Knox, A. G. and Walters, M. P. (1994). Extinct and endangered birds in the collections of the Natural History Museum. British Ornithologists’ Club Occasional Publications No. 1. Knystautas, A. J. V. and Sibnev, J. B. (1987). Die Vogelwelt Ussuriens. A. Ziemsen Verlag, Lutherstadt. Koelz, W. (1954). Ornithological studies. I. New birds from Iran, Afghanistan, and India. Contributions from the Institute for Regional Exploration, Ann Arbor, 1, 1–32. Koenig, A. (1911). Die Ergebnisse meiner Reise nach dem Sudan im Frühjahr 1910. Verhandlung des V. Internationalen Ornithologen-Kongresses in Berlin 30. Mai bis 4. Juni 1910, 469–545. Koford, R. R., Bowen, B. S., and Vehrencamp. S. L. (1986). Habitat saturation in communally-nesting Groove-billed Anis. American Naturalist, 127, 317–37. Koford, R. R., Bowen, B. S., and Vehrencamp, S. L. (1990). Groove-billed Anis: joint-nesting in a tropical cuckoo. In Cooperative breeding in birds, (eds. P. B. Stacey and W. D. Koenig), pp. 333–55. Cambridge University Press, Cambridge, U. K. Kokko, H., Johnstone, R. A., and Clutton-Brock, T. H. (2000). The evolution of cooperative breeding through group augmentation. Proceedings of the Royal Society of London, Series B, 268, 187–96.

Komdeur, J. (1994). The effect of kinship on helping in the cooperatively breeding Seychelles Warbler (Acrocephalus sechellensis). Proceedings of the Royal Society of London, Series B, 256, 47–52. Komdeur, J. (1996). Influence of helping and breeding experience on reproductive performance in the Seychelles Warbler: a translocation experiment. Behavioral Ecology, 7, 326–33. Kornerup, A. and Wanscher, J. H. (1967). Methuen handbook of colour. 2nd edn. Methuen, London. Köster, F. (1971). Zum Nistverhalten des Ani, Crotophaga ani. Bonner Zoologische Beiträge , 22, 4–27. Köster, F. (1976). Über die Höhenanpassung von Crotophaga ani und Tyrannus melancholicus in den Anden Kolumbiens. Journal für Ornithologie , 117, 73–99. Kotagama, S. and Fernando, P. (1994). A field guide to the birds of Sri Lanka. Wildlife Heritage Trust of Sri Lanka, Colombo. Kratter, A. W., Steadman, D. W., Smith, C. E., Filardi, C. E., and Webb, H. P. (2001). Avifauna of a lowland forest site on Isabel, Solomon Islands. Auk, 118, 472–83. Krause, D. W., Rogers, R. R., Forster, C. A., Hartman, J. H., Buckley, G. A., and Sampson, S. D. (1999). The Late Cretaceous vertebrate fauna of Madagascar: implications for Gondwanan palaeobiogeography. GSA Today, 9, 1–7. Krishnan, M. (1952). Koels (Eudynamys scolopacea) eating the poisonous fruit of the yellow oleander (Thevetia neriifolia). Journal of the Bombay Natural History Society, 50, 943–5. Kudo, H. and Dunbar, R. I. M. (2001). Neocortex size and social network size in primates. Animal Behaviour, 62, 711–22. Kunkel, P. (1974). Mating systems of tropical birds: the effects of weakness or absence of external reproductiontiming factors with special reference to prolonged pair bonds. Zeitschrift für Tierpsychologie , 34, 265–307. Kunz, G. H. (1850). Beitrag zur Naturgeschichte des europäischen Kukuks (Cuculus canorus L.). Naumannia, Archiv fur die Ornithologie. Deutschen OrnithologischeGesellschaft, Stuttgart, 1, 51–3. Kuroda, N. (1925). A contribution to the knowledge of the avifauna of the Riu kiu Islands and the vicinity. Kuroda, Tokyo.

Bibliography 563 Kuroda, N. (1936). Birds of the island of Java, vol. 2, Nonpasseres. Kuroda, Tokyo.

Lambert, F. and Yong, D. (1989). Some recent bird observations from Halmahera. Kukila, 4, 30–3.

Lachenaud, O. (2003). On the plumages of Senegal Coucal Centropus senegalensis and a putative observation of Black-throated Coucal C. leucogaster in Niger. Malimbus, 25, 55–6.

Land, H. C. (1970). Birds of Guatemala. Livingstone, Wynnewood, Pennsylvania.

Lack, D. (1954). The natural regulation of animal numbers. Clarendon, London.

Lang, M. R. and Wells, J. W. (1987). A review of eggshell pigmentation. World’s Poultry Science Journal, 43, 238–46.

Lack, D. (1968). Ecological adaptations for breeding in birds. Methuen, London.

Langmore, N. E., Hunt, S., and Kilner, R. M. (2003). Escalation of a coevolutionary arms race through host rejection of brood parasitic young. Nature, 422, 157–60.

Lack, D. (1976). Island biology illustrated by the land birds of Jamaica. Blackwell, Oxford.

Langrand, O. (1990). Guide to the birds of Madagascar. Yale University Press, New Haven.

Lahti, D. C. and Lahti, A. R. (2002). How precise is egg discrimination in weaverbirds? Animal Behaviour, 1135–42.

Langrand, O. and Sinclair, J. C. (1994). Additions and supplements to the Madagascar avifauna. Ostrich, 65, 302–10.

Laman, T. G., Gaither, J. C., and Lukas, D. E. (1996). Rain forest bird diversity in Gunung Palung National Park, West Kalimantan, Indonesia. Tropical Biodiversity, 3, 281–96. Laman, T. G., Burnaford, J. L., and Knott, C. D. (1997). Sunda Ground-cuckoo observations in Gunung Palung National Park, West Kalimantan. Kukila, 9, 183–5. Lamarche, B. (1980). Liste commentee des oiseaux du Mali, 1ère partie: Non-passereaux. Malimbus, 2, 121–59. Lamba, B. S. (1963). The nidification of some common Indian birds—part 1. Journal of the Bombay Natural History Society, 60, 121–33. Lamba, B. S. (1969). The nidification of some common Indian birds—part 12. Journal of the Bombay Natural History Society, 66, 72–80. Lamba, B. S. (1976). The Indian crows: a contribution to their breeding biology, with notes on brood parasitism on them by the Indian Koel. Records of the Zoological Survey of India, 71, 183–300. Lambert, F. R. (1989). Fig-eating by birds in a Malaysian tropical lowland rain forest. Journal of Tropical Ecology, 5, 401–12. Lambert, F. (1991). The conservation of fig-eating birds in Malaysia. Biological Conservation, 58, 31–40. Lambert, F. R. (1994). Notes on the avifauna of Bacan, Kasiruta and Obi, North Moluccas. Kukila, 7, 1–9. Lambert, F. R. and Marshall, A. G. (1991). Keystone characteristics of bird-dispersed Ficus in a Malaysian lowland rain forest. Journal of Ecology, 79, 793–809.

Langridge, H. P. (1990). Courtship feeding behavior in the Mangrove Cuckoo (Coccyzus minor). Florida Field Naturalist, 18, 55–6. Lank, D. B., Cooch, E. G., Rockwell, R. F., and Cooke, F. (1989). Environmental and demographic correlates of intraspecific nest parasitism in Lesser Snow Geese Chen caerulescens caerulescens. Journal of Animal Ecology, 58, 29–45. Lansdown, P. (1995). Ages of Great Spotted Cuckoos in Britain and Ireland. British Birds, 88, 141–9. Lanyon, S. M. and Omland, K. E. (1999). A molecular phylogeny of the blackbirds (Icteridae): five lineages revealed by cytochrome-b sequence data. Auk, 116, 629–39. Larkins, D. (1994a). The Channel-billed Cuckoo in far west Queensland: some records from pastoral literature. Australian Birds, 28, 5–7. Larkins, D. (1994b). The Channel-billed Cuckoo: behaviour at nests of Pied Currawongs. Australian Birds, 28, 7–10. Larson, L. M. (1930). Osteology of the California Road-runner Recent and Pleistocene. University of California Publications in Zoology, 32, 409–28. Lasiewski, R. C., Bernstein, M. H., and Ohmart, R. D. (1971). Cutaneous water loss in the Roadrunner and the Poor-will. Condor, 73, 470–2. La Touche, J. D. D. (1931–34). A handbook of the birds of eastern China, vol. 2. Taylor and Francis, London.

564 Bibliography Latta, S. C. and Wunderle, J. M. (1996). The composition and foraging ecology of mixed-species flocks in pine forest of Hispaniola. Condor, 98, 595–607. Latter, O. H. (1902). The egg of Cuculus canorus. An enquiry into the dimensions of the Cuckoo’s egg and the relation of the variations to the size of the eggs of the foster-parent, with notes on coloration, etc. Biometrika, 1, 164–76. Latter, O. H. (1905). The egg of Cuculus canorus. An attempt to ascertain from the dimensions of the Cuckoo’s egg if the species is tending to break up into subspecies, each exhibiting a preference for some foster parent. Biometrika, 4, 363–73. Lau, P., Bosque, C., and Strahl, S. D. (1998). Nest predation in relation to nest placement in the Greater Ani (Crotophaga major). Ornitologia Neotropical, 9, 87–92. Laubmann, A. (1939). Die Vögel von Paraguay,Vol. 1. Wissenschaftliche Ergebnisse der Deutschen Gran-ChacoExpedition. Strecker and Schröder, Stuttgart. Lavauden, L. (1932). Étude d’une petite collection d’oiseaux de Madagascar. Bulletin du Muséum National d’Histoire Naturelle, 2 (4), 629–40. Lavauden, L. (1937). Histoire physique, naturelle et politique de Madagascar. vol. 12, Oiseaux. Supplement to A. Milne Edwards & A. Grandidier (1879–1885). Société d’ Éditions Géographiques, Maritimes et Coloniales, Paris. Lawes, M. J. and Kirkman, S. (1996). Egg recognition and interspecific brood parasitism rates in Red Bishops (Aves: Ploceidae). Animal Behaviour, 52, 553–63.

Leach, H. A. C. (1929). Honeyeaters and cuckoos. Emu, 28, 177–82. LeCroy, M. and Peckover, W. S. (1983). Birds of the Kimbe Bay area, west New Britain, Papua New Guinea. Condor, 85, 297–304. LeCroy, M. and Peckover, W. S. (1998). Misima’s missing birds. Bulletin of the British Ornithologists’ Club, 118, 217–38. Lee, T.-Y. and Lawver, L. A. (1995). Cenozoic plate reconstruction of Southeast Asia. Tectonophysics, 152, 85–138. Legge, W.V. (1880). A history of the birds of Ceylon. The author, London. Leigh, E. G. (1999). Tropical forest ecology: a view from Barro Colorado Island. Oxford University Press, New York. Lekagul, B. and Round, P. D. (1991). A guide to the birds of Thailand. Saha Karn Bhaet, Bangkok. Leonard, P. M. (1998a). Concentration of Emerald Cuckoos Chrysococcyx cupreus. Zambia Bird Report 1997, 57. Leonard, P. M. (1998b). 1997 spring records. Zambia Bird Report 1997, 59–139. Léotaud, A. (1866). Oiseaux de l’île de la Trinidad, (Antilles). Chronicle Publishing Office, Port d’Espagne. Lerkelund, H. E., Moksnes, A., Røskaft, E., and Ringsby, T. H. (1993). An experimental test of optimal clutch size of the Fieldfare; with a discussion on why brood parasites remove eggs when they parasitize a host species. Ornis Scandinavica, 24, 95–102.

Lawrence, G. N. (1864). Catalogue of birds collected on Sombrero, W. I., with observations by A. A. Julien. Annals of the Lycee of Natural History of New York, 8, 41–6.

Levaillant, W. V. (1790). Travels from the Cape of GoodHope into the interior parts of Africa, vol. 2. Translated from the French by Elizabeth Helme, for William Lane, London (Johnson Reprint Corp., New York, 1972).

Lawson, W. J. (1962). Variation in the South African populations of the Coucal Centropus superciliosus Hemprich and Ehrenberg. Ostrich, 33, 45–7.

Levaillant, W.V. (1806). Histoire naturelle des oiseaux d’Afrique, vol. 5. Delachaussée, Paris.

Layard, E. L. (1874). Birds of South Africa. Bernard Quaritch, London. Laycock, G. (1985). When the paisano nests, lizards should lie low. Audubon, 87 (1), 80–3. Laymon, S. A. (1980). Feeding and nesting behavior of the Yellow-billed Cuckoo in the Sacramento Valley. Wildlife Management Administration Report, 80–2 , California Department of Fish and Wildlife, Sacramento.

Lévêque, R. (1968). Über Verbreitung, Bestandesvermehrung und Zug des Häherkuckucks Clamator glandarius (L.). in Westeuropa. Ornithologische Beobachter, 65, 43–71. Leverkühn, P. (1894). Über das Brutgeschäft der Crotophagiden. Journal für Ornithologie , 42, 44–80. Levings, S. C. and Windsor, D. M. (1982). Seasonal and annual variation in litter arthropod populations. In The ecology of a tropical forest: seasonal rhythms and long-term changes, (eds. E. G. Leigh, A. S. Rand and D. M.

Bibliography 565 Windsor), pp. 355–87. Smithsonian Institution Press, Washington, D.C. Levy, Mrs. C. (1984). Notes on the nestling of the Jamaican Lizard Cuckoo. Gosse Bird Club Broadsheet, 42, 6–8.

Liversidge, R. (1955). Observations on a Piet-my-vrou (Cuculus solitarius) and its host the Cape Robin (Cossypha caffra). Ostrich, 26, 18–27. Liversidge, R. (1961). Pre-incubation development of Clamator jacobinus. Ibis, 103A, 624.

Lewis, A. and Pomeroy, D. (1989). A bird atlas of Kenya. A. A. Balkema, Rotterdam.

Liversidge, R. (1971). The biology of the Jacobin Cuckoo Clamator jacobinus. Ostrich Suppl., 8, 117–37.

Lewis, F. (1898). Field-notes on the land-birds of Sabaragamuwa Province, Ceylon. Ibis, Ser. 7, vol. 4, 334–56, 524–51.

Livesey, T. R. (1933). Cuckoos in the Southern Shan States. Journal of the Bombay Natural History Society, 36, 997.

Lewthwaite, R. (1995). Blue Magpie as a foster parent of Koel. Hong Kong Bird Report, 1994, 222.

Livesey, T. R. (1936). Cuckoos in the Southern Shan States. Journal of the Bombay Natural History Society, 38, 401–3.

Li, G. (1991). [New records of bird from Sichuan.] Sichuan Journal of Zoology, 10, 27. [in Chinese] Ligon, J. D. (1993). The role of phylogenetic history in the evolution of contemporary avian mating and parental care systems. Current Ornithology, 10, 1–46. Ligon, J. D. (1997). A single functional testis as a unique proximate mechanism promoting sex-role reversal in coucals. Auk, 114, 800–1. Ligon, J. D. (1999). The evolution of avian breeding systems. Oxford University Press, Oxford. Lindholm, A. K. (1999). Brood parasitism by the Cuckoo on patchy Reed Warbler populations in Britain. Journal of Animal Ecology, 68, 293–309. Lindholm, A. K. (2000). Tests of phenotypic plasticity in Reed Warbler defences against cuckoo parasitism. Behaviour, 137, 43–60. Lindholm, A. and Lindén, A. (2003). Oriental Cuckoo in Finland. Alula, 4, 122–33. Lindholm, A. K. and Thomas, R. J. (2000). Differences between populations of Reed Warblers in defences against brood parasitism. Behaviour, 137, 25–42. Lindholm, A. K., Veneter, G. J., and Ueckermann, E. A. (1998). Persistence of passerine ectoparasites on the Diederik Cuckoo Chrysococcyx caprius. Journal of Zoology, 244, 145–54.

Livesey, T. R. (1938a). Egg-laying of the Khasia Hills Cuckoo (C. c. bakeri) in the nest of the Burmese Stone Chat (Saxicola caprata burmanica). Journal of the Bombay Natural History Society, 40, 125–7. Livesey, T. R. (1938b). Egg stealing by the Khasia Hills Cuckoo (Cuculus canorus bakeri Hartert). Journal of the Bombay Natural History Society, 40, 561–4. Livesey, T. R. (1938c). Call notes of the Burmese Plaintive Cuckoo (Cacomantis merulinus querulus). Journal of the Bombay Natural History Society, 40, 564–8. Livezey, B. C. (1993). An ecomorphological review of the Dodo (Raphus cucullatus) and the Solitaire (Pezophaps solitaria), flightless Columbiformes of the Mascarene Islands. Journal of Zoology, London, 130, 247–92. Livezey, B. C. (1997). A phylogenetic classification of waterfowl (Aves: Anseriformes), including selected fossil species. Annals of Carnegie Museum, 66, 457–96. Livezey, B. C. (1998). A phylogenetic analysis of the Gruiformes (Aves) based on morphological characters, with an emphasis on the rails (Rallidae). Philosophical Transactions of the Royal Society of London B, 353, 2077–151. Livezey, B. C. and Zusi, R. L. (2001). Higher-order phylogenetics of modern Aves based on comparative anatomy. Netherlands Journal of Zoology, 51, 179–205.

Linnaeus, C. (1758). Systema naturae. Regnum Animale (10th ed.). L. Salvii, Holmiae.

Lobb, M. G. (1983). Didric Cuckoo Chrysococcyx caprius in Cyprus—new to the Palaearctic. Bulletin of the British Ornithologists’ Club, 103, 111–2.

Linnaeus, C. (1766). Systema naturae. Regnum Animale (12th ed.). L. Salvii, Holmiae.

Loetscher, F. W. (1952). Striped Cuckoo fed by Rufousand-white Wren in Panama. Condor, 54, 169.

Linsley, M. D., Jones, M. J., and Marsden, S. J. (1998). A review of the Sumba avifauna. Kukila, 10, 60–90.

Loflin, R. K. (1982). Ani male apparently killed by other anis while attempting to parasitize nest. Auk, 99, 787–8.

566 Bibliography Loflin, R. K. (1983). Communal behaviors of the Smoothbilled Ani (Crotophaga ani). PhD dissertation, University of Miami, Coral Gables, Florida. Löhrl, H. (1979). Untersuchungen am Kuckuck Cuculus canorus (Biologie, Ethologie und Morphologie). Journal für Ornithologie , 120, 139–73. Loke, W. T. (1953). Some breeding birds of Singapore. Journal of the Bombay Natural History Society, 51, 590–6. Long, A. J. and Collar, N. J. (2002). Distribution, status and natural history of the Bornean Ground-cuckoo Carpococcyx radiatus. Forktail, 18, 111–9. López-Lanús, B., Berg, K. S., Stewe, R., and Salaman, P. G. W. (1999). The ecology and vocalizations of Banded Ground-cuckoo Neomorphus radiolosus. Cotinga, 11, 42–5. López-Lanús, B. (1999). Red-billed Ground-cuckoo Neomorphus pucheranii: hypothetical record in Ecuador. Cotinga, 12, 73. Lorber, P. (1984). Unrecorded display of the Striped Cuckoo. Honeyguide, 30, 76–7. Lorber, P. (1985). Dideric Cuckoo killed by African Masked Weaver. Honeyguide, 31, 55. Lord, E. A. R. (1956). The birds of Murphy’s Creek district, southern Queensland. Emu, 56, 100–28. Lorenz, K. (1935). Der Kumpan in der Umwelt des Vogels. Journal für Ornithologie , 83, 137–213, 289–413. Lorenzana, J. C. and Sealy, S. G. (2001). Fitness costs and benefits of cowbird egg ejection by Gray Catbirds. Behavioral Ecology, 12, 325–9. Lotem, A. and Nakamura, H. (1998). Evolutionary equilibrium in avian brood parasitism: an alternative for the ‘arms race-evolutionary lag’ concept. In Parasitic birds and their hosts, (eds. S. I. Rothstein and S. K. Robinson), pp. 223–35. Oxford University Press, Oxford. Lotem, A., Nakamura, H., and Zahavi, A. (1992). Rejection of cuckoo eggs in relation to host age: a possible evolutionary equilibrium. Behavioral Ecology, 3, 128–32. Lotem, A., Nakamura, H., and Zahavi, A. (1995). Constraints on egg discrimination and cuckoo-host coevolution. Animal Behaviour, 49, 1185–209. Louette, M. (1981a). The birds of Cameroon. An annotated check-list. Verhandelingen van de Koninklijke Academie voor Wetenschappen, Letteren en Schone Kunste van België, 43 (163), 1–295.

Louette, M. (1981b). Contribution to the ornithology of Liberia (Part 5). Revue de Zoologie Africaine, 95, 342–55. Louette, M. (1986). Geographical contacts between the taxa of Centropus in Zaïre, with the description of a new race. Bulletin of the British Ornithologists’ Club, 106, 126–33. Louette, M. (1989). Additions and corrections to the avifauna of Zaire (4). Bulletin of the British Ornithologists’ Club, 109, 217–25. Louette, M. and Herroelen, P. (1993). Status of migratory Cuculus cuckoos in Zaïre. Bulletin of the British Ornithologists’ Club, 113, 147–52. Louette, M. and Herroelen, P. (1994). A revised key for Cercococcyx cuckoos, taxonomic status of C. montanus patulus and its occurrence in Zaïre. Bulletin of the British Ornithologists’ Club, 114, 144–9. Lowe, C. M. and Minde, D. S. (1969). Thermoregulation in desert populations of roadrunners and doves. In Physiological systems in semiarid environments, (eds. C. C. Hoff and M. L. Âiesdesel). University of New Mexico Press, Albuquerque. Lowery, G. H. and Dalquest, W. W. (1951). Birds from the state of Veracruz, Mexico. University of Kansas Publications Museum of Natural History, 3 (4), 531–649. Loyn, R. H. (1980). Bird populations in a mixed eucalypt forest used for production of wood in Gippsland, Victoria. Emu, 80, 145–56. de Lucca, E. J. (1974). Cariotipos de 14 especes de Aves das Ordens Cuculiformes, Galliformes, Passeriformes e Tinamiformes. Revista Brasiliera de Pesquisas Móedicos e Biológias , 7, 253–63. Ludlow, F. and Kinnear, N. B. (1937). The birds of Bhutan and adjacent territories of Sikkim and Tibet. Ibis, Ser. 14, vol. 1, 1–46, 249–93, 467–504. Ludlow, F. and Kinnear, N. B. (1944). The birds of southeastern Tibet. Ibis, 86, 43–86, 176–208, 348–89. Lushington, C. (1949). Change in habits of the Ceylon Hawk-cuckoo (Hierococcyx varius ciceliae ( Phillips). Journal of the Bombay Natural History Society, 48, 582–4. Lynes, H. (1925). On the birds of north and central Darfur, part IV. Ibis, Ser. 12, vol. 1, 344–416. Lynes, H. (1938). Contribution to the ornithology of the southern Congo basin. Lynes-Vincent tour of 1933–34. Revue de Zoologie et de Botanique Africaines, 31, 1–128.

Bibliography 567 Lyon, B. E. and Eadie, J. M. (1991). Mode of development and interspecific avian brood parasitism. Behavioral Ecology, 2, 309–18.

MacKinnon, J. and Phillipps, K. (1993). The birds of Borneo, Sumatra, Java, and Bali, the Greater Sunda Islands. Oxford University Press, Oxford.

Macdonald, K. C. (1906). A list of birds found in the Myingyan district of Burma, part 2. Journal of the Bombay Natural History Society, 17, 492–504.

MacKinnon, J. and Phillipps, K. (2000). A field guide to the birds of China. Oxford University Press, Oxford.

MacDonald, M. (1957). Birds in my Indian garden. Jonathan Cape, London.

MacKinnon, K., Hatta, G., Halim, H., and Mangalik, A. (1996). The ecology of Kalimantan. Periplus Editions, Jakarta.

Macdonald, M. A. (1980). Observations on the Diederik Cuckoo in southern Ghana. Ostrich, 51, 75–9.

Mackness, B. (1979). The ecology of the Pheasant Coucal Centropus phasianinus (Latham) in Australia. Sunbird, 10, 1–8.

Macedo, R. H. (1991). Communal breeding and social organization of the Guira Cuckoo (Guira guira) in central Brazil. PhD dissertation, University of Oklahoma, Norman.

Mackworth-Praed, C. W and Grant, C. H. B. (1970). Birds of west central and western Africa. African Handbook of Birds, Series III, vol. 1. Longman, London.

Macedo, R. H. (1992). Reproductive patterns and social organization of the communal Guira Cuckoo (Guira guira) in central Brazil. Auk, 109, 786–99.

Maclean, G. L. (1993). Roberts’ birds of southern Africa. (6th edn.). Trustees of the John Voelcker Bird Book Fund, Cape Town.

Macedo, R. H. (1994). Inequities in parental effort and costs of communal breeding in the Guira Cuckoo. Ornitologia Neotropical, 5, 79–90.

Macleod, J. G. R. and Hallack, M. (1956). Some notes on the breeding of Klaas’s Cuckoo. Ostrich, 27, 2–5.

Macedo, R. H. and Bianchi, C. A. (1997a). Communal breeding in tropical Guira Cuckoos Guira guira: sociality in the absence of a saturated habitat. Journal of Avian Biology, 28, 207–15.

Madoc, G. C. (1956a). Observations on two Malayan cuckoos. Malayan Nature Journal, 10, 97–103. Madoc, G. C. (1956b). A further note on the Lesser Cuckoo. Malayan Nature Journal, 10, 134–8.

Macedo, R. H. and Bianchi, C. A. (1997b). When birds go bad: circumstantial evidence for infanticide in the communal South-American Guira Cuckoo. Ethology, Ecology and Evolution, 9, 45–54.

Madroño N., A., Clay, R. P., Robbins, M. B., Rice, N. H., Faucett, R. C., and Lowen, FJ. C. (1997). An avifaunal survey of the vanishing interior Atlantic forest of San Rafael National Park, Departments Itapúa/Caazapá, Paraguay. Cotinga, 7, 45–53.

Macedo, R. H. and Melo, C. (1999). Confirmation of infanticide in the communally breeding Guira Cuckoo. Auk, 116, 847–51.

Magalhães, J. C. Reis de. (1999). As aves na fazenda Barreiro Rico. Plêiade, São Paulo.

Macedo, R. H. F., Cariello, M. O., Graves, J., and Schwabl, H. (2004). Reproductive partitioning in communally breeding Guira Cuckoos, Guira guira. Behavioral Ecology and Sociobiology, 55, 213–22.

Magrath, R. D. and Whittingham, L. A. (1997). Subordinate males are more likely to help if unrelated to the breeding female in cooperatively breeding White-browed Scrub-wrens. Behavioral Ecology and Sociobiology, 41, 185–92.

MacGillivray, W. (1914). Notes on some North Queensland birds. Emu, 13, 132–86. Machado, A. B. M., Fonseca, G. A. B., Machado, R. B., Aguillar, L. M. de S., and Lins, L. V. (eds.). (1998). Livro vermelho das espécies ameaçadas de extinção da fauna de Minas Gerais. Belo Horizonte, Fundação Biodiversitas, São Paulo. Mackenzie, J. M. D. (1918). Some further notes on cuckoos in Maymyo. Journal of the Bombay Natural History Society, 25, 742–5.

Mahabal, A. and Lamba, B. S. (1987). On the birds of Poona and vicinity. Zoological Survey of India Miscellaneous Publication Occasional Paper, 94. Makatsch, W. (1937). Der Brutparasitismus der Kuckucksvögel . Neumann Verlag, Berlin. Makatsch, W. (1955). Der Brutparasitismus in der Vogelwelt. Neumann Verlag, Radebeul and Berlin. Makatsch, W. (1971). Einige Bemerkungen über die parasitären Kuckucke. Zoologische Abhandlungen

568 Bibliography Staatliches Museum für Tierkunde in Dresden , 30, 247–83. Malchevsky, A. S. (1960). On the biological races of the Common Cuckoo Cuculus canorus L. in the territory of the European part of the USSR. XII International Ornithological Congress, Helsinki, 1958, 464–70. Malchevsky, A. S. (1987). [A Cuckoo and her offspring]. Leningrad University, Leningrad. (in Russian). Maller, C. J. and Jones, D. N. (2001). Vocal behaviour of the Common Koel, Eudynamys scolopacea, and implications for mating systems. Emu, 101, 105–12. Marceliano, M. L. Videira. (1996). Estudo osteólogico e miológico do cranio de Opisthocomus hoazin (Müller, 1776) (Aves: Opisthocomidae), comparado com algumas espécies de Cracidae, Musophagidae e Cuculidae. Boletim do Museu Paraense Emilio Goeldi, Série Zoologia, Belém, Pará, 12 (2), 95–246. Marchant, S. (1942). Some birds of the Owerri Province, S. Nigeria. Ibis, 1942, 137–96. Marchant, S. (1960). The breeding of some southwestern Ecuadorian birds. Ibis, 102, 349–82, 584–99. Marchant, S. (1972). Evolution of the genus Chrysococcyx. Ibis, 114, 219–33. Marchant, S. (1986). A note on brood parasitism by the Shining Bronze-cuckoo. Australian Birds, 20, 82–6. Marchant, S. and Höhn, E. O. (1980). Field notes on the Fan-tailed Cuckoo. Emu, 80, 77–80. Marchetti, K. (1992). Costs to host defence and the persistence of parasitic cuckoos. Proceedings of the Royal Society of London, Series B, 248, 41–5. Marchetti, K., Nakamura, H., and Gibbs, H. L. (1998). Host-race formation in the Common Cuckoo. Science, 282, 471–2. Marcodes, B. and Rinke, D. (2000). Der HaubenSeidenkuckuck. Gefiederte Welt, 124, 189–91. Marks, B. D., Hackett, S. J., and Capparella, A. P. (2002). Historical relationships among Neotropical lowland forest areas of endemism as determined by mitochondrial DNA sequence variation within the Wedge billed Woodcreeper (Aves: Dendrocolaptidae: Glyphorynchus spurius). Molecular Phylogenetics and Evolution, 24, 153–67. Markus, M. B. (1964). Egg polymorphism in the Diederik Cuckoo (Chrysococcyx caprius (Boddaert)) at Pretoria. Ostrich, 35, 123. Marler, P. R. and Hamilton, W. J., III. (1966). Mechanisms of animal behavior. John Wiley, New York.

Marsh, O. C. (1872). Notice of some new Tertiary and Post-Tertiary birds. American Journal of Science and Arts, 4, 256–62. Marshall, J. T. (1957). Birds of pine-oak woodland in southern Arizona and adjacent Mexico. Pacific Coast Avifauna 32. Martens, J. and Eck, S. (1995). Towards an ornithology of the Himalayas: systematics, ecology and vocalizations of Nepal birds. Bonner Zoologische Monographien, 38. Martin, L. D. and Mengel, R. M. (1984). A new cuckoo and a chachalaca from the early Miocene of Colorado. Papers in vertebrate paleontology honoring Robert Warren Wilson, (ed. R. M. Mengel). Special Publications of the Carnegie Museum of Natural History, 9, 171–7. Martínez, J. G., Burke, T., Dawson, D., Soler, J. J., Soler, M., and Møller, A. P. (1998a). Microsatellite typing reveals mating patterns in the brood parasitic Great Spotted Cuckoo (Clamator glandarius). Molecular Ecology, 7, 289–97. Martínez, J. G., Soler, J. J., Soler, M., and Burke, T. (1998b). Spatial patterns of egg laying and multiple parasitism in a brood parasite: a non-territorial system in the Great Spotted Cuckoo (Clamator glandarius). Oecologia, 117, 286–94. Martins, F. de C. and Donatelli, R. J. (2001). Estratégia alimentar de Guira guira (Cuculidae, Crotophaginae) na região centro-oeste do Estado de São Paulo. Ararajuba, 9, 89–94. Mason, I. J. (1982). The identity of certain early Australian types referred to the Cuculidae. Bulletin of the British Ornithologists’ Club, 102, 99–106. Mason, I. J. (1997a). Cuculidae, Centropodidae. In Aves (Columbidae to Coraciidae). Zoological Catalogue of Australia, 37 (2), (eds. W. W. K. Houston and A. Wells), pp. 219–63. CSIRO, Melbourne. Mason, I. J. (1997b). Centropodidae. In Aves (Columbidae to Coraciidae). Zoological Catalogue of Australia, 37 (2), (eds. W. W. K. Houston and A. Wells), pp. 255–63. CSIRO, Melbourne. Mason, I. J. and Forrester, R. I. (1996). Geographical differentiation in the Channel-billed Cuckoo Scythrops novaehollandiae Latham, with description of two new subspecies from Sulawesi and the Bismarck Archipelago. Emu, 96, 217–33. Mason, I. J., McKean, J. L., and Dudzinski, M. L. (1984). Geographical variation in the Pheasant Coucal Centropus phasianinus (Latham) and a description of a new subspecies from Timor. Emu, 84, 1–15.

Bibliography 569 Mason, P. (1985). The nesting biology of some passerines of Buenos Aires, Argentina. Ornithological Monographs, 36, 954–72.

Mayr, E. (1938a). Notes on a collection of birds from south Borneo. Bulletin of the Raffles Museum, Singapore, Straits Settlement, no. 11, 5–46.

Masterson, A. N. B. and Parkes, D. A. (1998). Green Coucal in the Middle Zambezi Valley. Honeyguide, 44, 89–90.

Mayr, E. (1938b). The birds of the Vernay-Hopwood Chindwin Expedition. Ibis, Ser. 14, vol. 2, 277–320.

Masuda, T. and Ramanampamonjy, J. R. (1996). Home range and nest site characteristics of three sympatric species of couas Coua ruficeps, C. coquereli and C. cristata in western Madagascar. In Social evolution of birds in Madagascar, with special respect to vangas, (ed. S. Yamagishi), pp. 93–7. Osaka City University, Osaka. Mathews, G. M. (1912). Notes on Australian cuckoos. Austral Avian Record, 1, 2–22. Mathews, G. M. (1918). The birds of Australia, vol. 7. Witherby et al., London. Matsuda, T. and Uchida, H. (1990). [Rearing strategies of Cuckoos.] Color Nature Series 68. Akane-shobo, Tokyo. (in Japanese). Mauersberger, G. (1980). Ökofaunistische und biologische Beiträge zur Avifauna Mongolica. II, Gruiformes bis Passeriformes. Mitteilungen aus der Zoologischen Museum in Berlin 56, Annalen für Ornithologie , 4, 77–164.

Mayr, E. (1939). The winter quarters of Chalcites malayanus minutillus (Gould). Emu, 39, 128–9. Mayr, E. (1941). List of New Guinea birds. American Museum of Natural History, New York. Mayr, E. (1942). Systematics and the origin of species. Columbia University Press, New York. Mayr, E. (1944a). Birds collected during the Whitney South Sea Expeditions, 54. Notes on some genera from the southwest Pacific. American Museum Novitates, 1269, 1–8. Mayr, E. (1944b). The birds of Timor and Sumba. Bulletin of the American Museum of Natural History, 83, 123–94. Mayr, E. (1944c). Wallace’s line in the light of recent zoogeographic studies. Quarterly Review of Biology, 19, 1–14. Mayr, E. (1963). Animal species and evolution. Oxford University Press, Oxford.

Maumary, L. and Duperrex, H. (1991). Le coucou-geai Clamator glandarius nicheur dans le sud Marocain. Alauda, 59, 53.

Mayr, E. and Diamond, J. M. (2001). The birds of northern Melanesia. Speciation, ecology, and biogeography. Oxford University Press, New York.

Maurício, G. N. and Dias, R. A. (2000). New distributional information for birds in southern Rio Grande do Sul, Brazil, and the first record of the Rufous Gnateater Conopophaga lineata for Uruguay. Bulletin of the British Ornithologists’ Club, 120, 230–7.

Mayr, E. and Gilliard, E. T. (1954). Birds of central New Guinea. Bulletin of the American Museum of Natural History, 103, 311–74.

Maxwell, P. H. (1938). Bernstein’s Ground-cuckoo (Centropus bernsteini). Avicultural Magazine, series 5, volume 3, 125. Mayhew, W. W. (1971). Desert encounter: a Roadrunner and a sidewinder. Fauna, 1, 17–9. Mayr, E. (1932). Birds collected during the Whitney South Sea Expedition, 19. Notes on the Bronze Cuckoo Chalcites lucidus and its subspecies. American Museum Novitates, 520, 1–9. Mayr, E. (1937). Birds collected during the Whitney South Sea Expedition, 35. Notes on New Guinea birds, 2. American Museum Novitates, 939, 1–14.

Mayr, E. and Meyer de Schauensee, R. (1940a). Zoological results of the Denison-Crockett expedition to the South Pacific for the Academy of Natural Sciences of Philadelphia 1937–1938. Part 1. The birds of the island of Biak. Proceedings of the Academy of Natural Sciences of Philadelphia, 91, 1–37. Mayr, E. and Meyer de Schauensee, R. (1940b). Zoological results of the Denison-Crockett expedition to the South Pacific for the Academy of Natural Sciences of Philadelphia 1937–1938. Part 4. Birds from northwest New Guinea. Proceedings of the Academy of Natural Sciences of Philadelphia, 91, 97–144. Mayr, E. and Meyer de Schauensee, R. (1940c). Zoological results of the Denison-Crockett expedition to the South Pacific for the Academy of Natural

570 Bibliography Sciences of Philadelphia 1937–1938. Part 5. Birds from the western Papuan islands. Proceedings of the Academy of Natural Sciences of Philadelphia, 91, 145–163.

McKitrick, M. C. (1991). Phylogenetic analysis of avian hindlimb musculature. Miscellaneous Publications of the University of Michigan Museum of Zoology, 179.

Mayr, E. and Rand, A. L. (1936). Results of the Archbold Expeditions. 10. Two new subspecies of birds from New Guinea. American Museum Novitates, 868, 1–3.

McLean, E. B. (1995). Smooth-billed Ani, a new Ohio bird. Ohio Cardinal, 19, 1–5.

Mayr, E. and Rand, A. L. (1937). Results of the Archbold Expeditions. 14. Results of the 1933–1934 Papuan Expedition. Bulletin of the American Museum of Natural History, 73, 1–248. Mayr, E. and Vaurie, C. (1948). Evolution in the family Dicruridae (birds). Evolution, 2, 238–65. Mayr, G. (1998a). “Coraciiforme” und “Piciforme” Kleinvögel aus dem Mittel-Eozän der Grube Messel (Hessen, Deutschland). Courier Forschungsinstitut Senckenberg, 205. Mayr, G. (1998b). A new family of Eocene zygodactyl birds. Senckenbergiana Lethaea, 78, 199–209. Mayr, G. (1999). A new trogon from the Middle Oligocene of Céreste, France. Auk, 116, 427–34. Mayr, G. (2000). Charadriiform birds from the Early Oligocene of Céreste (France) and the Middle Eocene of Messel (Hessen, Germany). Geobios, 33, 625–36. McBride, H. C. A. (1984). Multiple feeding of young Cuckoo. British Birds, 77, 422–3. McCaskie, G., DeBenedictis, P., Erickson, R., and Morlan, J. (1988). Birds of northern California. Golden Gate Audubon Society, Berkeley. McClure, H. E. (1998). Migration and survival of the birds of Asia. White Lotus Press, Bangkok, Thailand. McGill, A. R. and Goddard, M. T. (1979). The Little Bronze Cuckoo in New South Wales. Australian Birds, 14, 23–4. McGilp, J. N. (1941). Habits of South Australian cuckoos. South Australian Ornithologist, 15, 115–24. McGregor, R. C. (1904). The birds of Calayan and Fuga, Babuyan Group. Bulletin of the Philippine Museum, 4, 1–34. McGregor, R. C. (1905). Birds from Mindoro and small adjacent islands. Bulletin of the Bureau of Government Laboratories, Manila, 34, 5–27. McGregor, R. C. (1909). A manual of Philippine birds. Manila, Bureau of Printing.

McLean, E. B., White, A. M., and Matson, T. O. (1995). Smooth-billed Ani (Crotophaga ani L.), a new species of bird for Ohio. Ohio Journal of Science, 95, 335–6. McLean, I. G. (1982). Whitehead breeding, and parasitism by Long-tailed Cuckoos. Notornis, 29, 156–8. McLean, I. G. (1988). Breeding behaviour of the Longtailed Cuckoo on Little Barrier Island. Notornis, 35, 89–98. McLean, I. G. and Griffin, J. M. (1991). Structure, function, and mimicry in begging calls of passerines and cuckoos. Acta XX Congressus Internationalis Ornithologicus, 2, 1273–84. McLean, I. G. and Waas, J. R. (1987). Do cuckoo chicks mimic the begging calls of their hosts? Animal Behaviour, 35, 1896–8. McLean, I. G., Gill, B. J., and Curthoys, L. P. (1986). Mortality, interference and injury at Whitehead nests. Notornis, 33, 266–8. McNair, D. B. (1991). Copulation in the Mangrove Cuckoo (Coccyzus minor). Florida Field Naturalist, 19, 84–5. McNair, D. B., Sibley, F., Massia, E. B., and Frost, M. D. (2002). Ground-based Nearctic-Neotropic landbird migration during autumn in the eastern Caribbean. In Studies in Trinidad and Tobago Ornithology Honouring Richard ffrench, (eds. F. E. Hayes and S. A. Temple), pp. 86–103. Department of Life Sciences, University of the West Indies, St Augustine, Occasional Papers, 11. McNeil, R. (1968). Biometric analysis of testicular bilateral symmetry of Crotophaga major. Ibis, 110, 87–9. McWhirter, D. W., Ikenaga, H., Iozawa, H., Shoyama, M., and Takehara, K. (1996). A check-list of the birds of Okinawa Prefecture with notes on recent status including hypothetical records. Bulletin of the Okinawa Prefectural Museum, 22, 33–152. Mearns, B. and Mearns, R. (1998). The bird collectors. Academic Press, London. Mears, A. and Oates, E. W. (1907). On the birds of the Chindwin, Upper Burma. Journal of the Bombay Natural History Society, 18, 78–87.

Bibliography 571 Medway, Lord and Wells, D. R. (1976). The birds of the Malay peninsula, vol. 5. H. F. & G. Witherby, London.

Merritt, J. H. (1951). Little Orphan Ani. Audubon, 53, 224–31.

Mees, G. F. (1949). Over de roeptijd van Cuculus micropterus concretus S. Müller. Limosa, 22, 331.

Merriwether, D. A., Friedlaender, J. S., Mediavilla, J., Mgone, C., Gentz, F., and Ferrell, R. E. (1999). Mitochondrial DNA variation is an indicator of Austronesian influence in island Melanesia. American Journal of Physical Anthropology, 110, 243–70.

Mees, G. F. (1964). Notes on two small collections of birds from New Guinea. Zoologische Verhandelingen, 66, 1–37. Mees, G. F. (1965). The avifauna of Misool. Nova Guinea, 31, 139–203. Mees, G. F. (1970). On some birds from southern Mexico. Zoologische Mededelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 44, 237–45. Mees, G. F. (1971). The Philippine subspecies of Centropus bengalensis (Gmelin) (Aves, Cuculidae). Zoologische Mededelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 45, 189–91. Mees, G. F. (1972). Die Vögel der Insel Gebe. Zoologische Mededelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 46, 69–89. Mees, G. F. (1979). Die Nachweise von Cuculus canorus L. im Indo-Australischen Raum. Mitteilungen aus der Zoologischen Museum in Berlin 55, Annalen für Ornithologie, 3, 127–34, Tafel 14. Mees, G. F. (1982). Birds from the lowlands of southern New Guinea (Merauke and Koembe). Zoologische Verhandelingen, Rijksmuseum van Natuurlijke Historie te Leiden 191. Mees, G. F. (1986). A list of the birds recorded from Bangka Island, Indonesia. Zoologische Verhandelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 232, 1–176. Mees, G. F. (1996). Geographical variation in birds of Java. Publications of the Nuttall Ornithological Club, 26. Meinertzhagen, R. (1954). Birds of Arabia. Oliver and Boyd, Edinburgh. Meinzer, W. (1993). The Roadrunner. Texas Tech University Press, Lubbock, Texas. Melo, C. and Macedo, R. H. F. (1997). Mortalidade em ninhadas de Guira guira (Cuculidae): competição por rercursos? Ararajuba, 5, 45–52. Mendelssohn, H. and Golani, I. (1986). Adaptability and limitations of a nest parasite. XIX Congressus Internationalis Ornithologicus, Ottawa, Abstract, 872.

Meyer, A. (1994). Shortcomings of the cytochrome b gene as a molecular marker. Trends in Ecology and Evolution, 9, 278–80. Meyer, A. B. (1879). Field-notes on the birds of Celebes. Part 1. Psittaci, Rapaces, and Picariae. Ibis, Ser. 4, vol. 3, 43–70. Meyer, O. (1927). Zur Lebensweise zweier Vogelarten des Bismarck-Archipels. Ornithologische Monatsberichte, 35, 139–40. Meyer, O. (1931). Ueber Nest und Gelege von Centropus violaceus. Ornithologische Monatsberichte, 39, 148. Meyer, O. (1933). Vogeleier und Nester aus Neubrittanien, Südsee. Beiträge zur Fortpflanzungsbiologie der Vögel , 9, 122–36, 182–5. Meyer, O. (1936). Die Vögel des Bismarckarchipel win hilfsbuch zur Erkennung der Vogelarten. Beobachtungen und Studien der Missionare vom Hlst. Herzen Jesu in der Südsee, Band 2 . Katholische Mission, Vunapope, Kokopo. Meyer de Schauensee, R. (1940). The birds of the Batu Islands. Proceedings of the Academy of Natural Sciences of Philadelphia, 92, 23–42. Meyer de Schauensee, R. (1957). On some avian types, principally Gould’s, in the collection of the Academy. Proceedings of the Academy of Natural Sciences of Philadelphia, 109, 123–246. Meyer de Schauensee, R. (1970). A guide to the birds of South America. Academy of Natural Sciences of Philadelphia, Philadelphia. Meyer de Schauensee, R. and Phelps, W. H., Jr. (1978). A guide to the birds of Venezuela. Princeton University Press, Princeton. Meyer de Schauensee, R. and Ripley, S. D. (1939). Zoological results of the George Vanderbilt Sumatran expedition 1936–1939. Part 1. Birds from Atjeh. Proceedings of the Academy of Natural Sciences of Philadelphia, 91, 311–68.

572 Bibliography Michaux, B. (1998). Terrestrial birds of the Indo-Pacific. In Biogeography and geological evolution of SE Asia, (eds. R. Hall and J. D. Holloway), pp. 361–391. Backhuys, Leiden, The Netherlands. Midya, S. and Brahmachary, R. L. (1991). The effect of birds upon germination of banyan (Ficus bengalensis) seeds. Journal of Tropical Ecology, 7, 537–8.

Milne-Edwards, A. and Grandidier, A. (1879). Histoire Physique, Naturelle et Politique de Madagascar, vol. 12. Histoire Naturelle des Oiseaux, Tome I. Paris. Milne-Edwards, A. and Grandidier, A. (1895). Sur des ossements d’oiseaux, vol. I.—texte. Imprimerie Nationale, Paris, 779 pp.

Mikhailov, K. E. (1997). Avian eggshells: an atlas of scanning electron micrographs. British Ornithologists’ Club Occasional Publications, no. 3.

Milon, P. (1950). Description d’une sous-espèce nouvelle d’oiseau de Madagascar. Bulletin du Muséum National d’Histoire Naturelle, Serie A, Zoologie, Paris, 2e série 22, 65–6.

Miksik, I., Honan, V., and Deyl, Z. (1996). Avian eggshell pigments and their variability. Comparative Biochemistry and Physiology, 113B, 606–12.

Milon, P. (1952). Notes sur le genre Coua. L’Oiseau et la Revue Française d’Ornithologie, 22, 75–90.

Miller, A. H. (1932). Observations on some breeding birds of El Salvador, Central America. Condor, 34, 8–17. Miller, A. H. (1955). The avifauna of the Sierra del Carmens of Coahuila, Mexico. Condor, 57, 154–78. Miller, A. H. (1963). Seasonal activity and ecology of the avifauna of an American equatorial cloud-forest. University of California Publications in Zoology, 66, 1–78. Miller, A. H. (1965). Capacity for photoperiodic response and endogenous factors in the reproductive cycles of an equatorial sparrow. Proceedings of the National Academy of Sciences USA, 54, 97–101. Miller, A. H. and Stebbins, R. C. (1964). The lives of desert animals in Joshua Tree National Monument. University of California Press, Berkeley. Miller, G. S. (1894). The ground cuckoo of Andros Island. Auk, 11, 164–5. Miller, L. (1943). The Pleistocene birds of San Josecito Cavern, Mexico. University of California Publications in Zoology, 47, 143–68. Mille-Horsin. (1921). Souvenirs d’un naturaliste en Afrique occidentale Française. A. Au Togo. 6. Le Coq de pagode ou Coucal a éperons (Centropus monachus Rupp.). L’Oiseau, 21, 75–9. Mills, K. (1987). The Pheasant Coucal in southern New South Wales, and a recent record from the Illwarra. Australian Birds, 21, 22–4. Milne-Edwards, A. (1892). Sur les oiseaux fossiles des dépots eocénes de phosphate de chaux du sud de la France. Second Congrès Ornithologique International, Budapest, 1891, 2, 60–80.

Milon, P. (1959). Sur la migration et la reproduction à Madagascar du Cuculus poliocephalus rochii. Ostrich Suppl., 3, 242–9. Milon, P., Petter, J.-J., and Randrianasolo, G. (1973). Faune de Madagascar, 35, Oiseaux. ORSTOM, Tananarive, and CNRS, Paris. Milson, J., Kaye, S., and Sardjono. (1996). Extension, collision and curvature in the eastern Banda arc. In Tectonic evolution of Southeast Asia (eds. R. Hall and D. J. Blundell), pp. 85–94. London, Geological Society Special Publication, 106. Mindell, D. P. (ed.). (1997). Avian molecular evolution and systematics. Academic Press, New York. Mindell, D. P., Sorenson, M. D., Huddleston, C. J., Miranda, H. C., Knight, A., Sawchuck, S. J., and Yuri, T. (1997). Phylogenetic relationships among and within select avian orders based on mitochondrial DNA. In Avian molecular evolution and systematics, (ed. D. P. Mindell), pp. 214–47. Academic Press, New York. Mindell, D. P., Sorenson, M. D. Dimcheff, D. E., Hasegawa, M., Ast, J. C., and Yuri, T. (1999). Interordinal relationships of birds and other reptiles based on whole mitochondrial genomes. Systematic Biology, 48, 138–52. Miranda, H. C., Kennedy, R. S., Sison, R. V., Gonzales, P. C., and Ebreo, M. F. (2000). New records of birds from the island of Panay, Philippines. Bulletin of the British Ornithologists’ Club, 120, 266–80. Mitchell, M. H. (1957). Observations on birds of southeastern Brazil. Toronto University Press, Toronto. Mitsch, H. (1974). Bemerkungswerte Verhaltensweisen bei den Ani. Gefiederte Welt, 98 (7), 124–7.

Bibliography 573 Mlíkovsky, J. (1989). Brain size in birds: 3. Columbiformes through Piciformes. Vestník Ceskoslovenské Spolecnosti Zoologické, 53, 252–64.

Moksnes, A., Røskaft, E., and Tysse, T. (1995). On the evolution of blue Cuckoo eggs in Europe. Journal of Avian Biology, 26, 13–9.

Mlodinow, S. T. and Karlson, K. T. (1999). Anis in the United States and Canada. North American Birds, 53, 237–45.

Moksnes, A., Røskaft, E., Hagen, L. G., Honza, M., Mørk, C., and Olsen, P. H. (2000). Common Cuckoo Cuculus canorus and host behaviour at Reed Warbler Acrocephalus scirpaceus nests. Ibis, 142, 247–58.

Mohan, D. (1997). Birds of New Forest, Dehra Dun, India. Forktail, 12, 19–30. Moksnes, A. and Røskaft, E. (1987). Cuckoo host interactions in Norwegian mountain areas. Ornis Scandinavica, 18, 168–72. Moksnes, A. and Røskaft, E. (1988). Responses of Fieldfares Turdus piliaris and Bramblings Fringilla montifringilla to experimental parasitism by the Cuckoo Cuculus canorus. Ibis, 130, 535–9. Moksnes, A. and Røskaft, E. (1989). Adaptations of Meadow Pipits to parasitism by the Common Cuckoo. Behavioral Ecology and Sociobiology, 24, 25–30. Moksnes, A. and Røskaft, E. (1992). Responses of some rare Cuckoo hosts to mimetic model Cuckoo eggs and to foreign conspecific eggs. Ornis Scandinavica, 23, 17–23. Moksnes, A. and Røskaft, E. (1995). Egg-morphs and host preferences in the Common Cuckoo (Cuculus canorus): an analysis of cuckoo and host eggs from European museums. Journal of Zoology, 236, 625–46. Moksnes, A., Røskaft, E., Braa, A. T., Korsnes, L., Lampe, H. M., and Pedersen, H. C. (1990). Behavioural responses of potential hosts towards artificial Cuckoo eggs and dummies. Behaviour, 116, 64–89. Moksnes, A., Røskaft, E., and Braa, A. T. (1991). Rejection behavior by Common Cuckoo hosts towards artificial brood parasite eggs. Auk, 108, 348–54. Moksnes, A., Røskaft, E., and Korsnes, L. (1993a). Rejection of Cuckoo (Cuculus canorus) eggs by Meadow Pipits (Anthus pratensis). Behavioral Ecology, 4, 120–7. Moksnes, A., Røskaft, E., Bicik, V. V., Honza, M., and Øien, I. J. (1993b). Cuckoo Cuculus canorus parasitism on Acrocephalus warblers in southern Moravia in the Czech Republic. Journal für Ornithologie , 134, 425–34. Moksnes, A., Røskaft, E., and Mørkvedt Solli, M. (1994). Documenting puncture ejection of parasitic eggs by Chaffinches Fringilla coelebs and Blackcaps Sylvia atricapilla. Fauna Norvegica Series C, Cinclus, 17, 115–8.

Molnár, B. (1944). The Cuckoo in the Hungarian plain. Aquila, 51, 100–12. Monk, K. A., de Fretes, Y., and Reksodiharjo-Lilley, G. (1997). The ecology of Nusa Tenggara and Maluku. The Ecology of Indonesia Series, Vol. 5. Periplus Editions, Singapore. Moore, A. (1984). Levaillant’s Cuckoo Clamator levaillantii fed by Brown Babbler Turdoides plebejus. Malimbus, 6, 94–5. Moore, R. T. (1934). A review of the races of Geococcyx velox. Transactions of the San Diego Society of Natural History, 7, 455–70. Moore, W. S. (1995). Inferring phylogenies from mtDNA variation: mitochondrial gene-trees versus nuclear-gene trees. Evolution, 49, 718–26. Moreau, R. E. (1937). Migrant birds in Tanganyika Territory. Tanganyika Notes and Records, 4, 1–34. Moreau, R. E. (1938). A contribution to the biology of the Musophagiformes, the so-called plantain-eaters. Ibis, Ser. 14, vol. 2, 639–71. Moreau, R. E. (1950). The breeding seasons of African birds. 1. Land birds. Ibis, 92, 223–67. Moreau, R. E. (1972). The Palaearctic-African bird migration systems. Academic Press, London. Moreau, R. E. and Chapin, J. P. (1951). The African Emerald Cuckoo, Chrysococcyx cupreus. Auk, 68, 174–89. Moreau, R. E. and Moreau, W. M. (1937). Biological and other notes on some East African birds. Ibis, Ser. 14, vol. 1, 152–74. Moreau, R. E. and Moreau, W. M. (1939). Observations on some East African birds. Ibis, Ser. 14, vol. 3, 296–323. Morel, G. J. and Morel, M.-Y. (1990). Les oiseaux de Sénégambie, notices et cartes de distribution . ORSTOM, Paris. Morel, M.-Y. (1973). Contribution à l’étude dynamique de la population de Lagonosticta senegala L.

574 Bibliography (Estrildides) à Richard-Toll (Sénégal). Interrelations avec le parasite Hypochera chalybeata (Müller) (Viduines). Mémoires du Muséum National d’Histoire Naturelle, serie A, Zoologie, 78. Morgan, G. S. (1994). Late Quaternary fossil vertebrates from the Cayman Islands. In The Cayman Islands: natural history and biogeography, (eds. M. A. Brunt and J. Davis), pp. 153–64. Kluwer Academic Publishers, Dordrecht, The Netherlands. Morgan, P. J. (1975). A catalogued specimen of Coua delalandei (Temminck) (Cuculidae) in Merseyside County Museums, Liverpool. Bulletin of the British Ornithologists’ Club, 95, 62–4. Morgensen, J. (1927). Nota sobre el parasitismo del “crespin” (Tapera naevia). El Hornero, 4, 68–70. Moritz, C., Dowling, T. E., and Brown, W. M. (1987). Evolution of animal mitochondrial DNA: relevance for population biology and systematics. Annual Review of Ecology and Systematics, 18, 269–92. Morris, P. and Hawkins, F. (1998). Birds of Madagascar, a photographic guide. Yale University Press, New Haven. Morrison, C. M. (1947). Field notes on some Gold Coast birds. Nigerian Field, 12, 59–64. Morton, E. S. and Farabaugh, S. M. (1979). Infanticide and other adaptations of the nestling Striped Cuckoo Tapera naevia. Ibis, 121, 212–3. Moskát, C. and Honza, M. (2002). European Cuckoo Cuculus canorus parasitism and host’s rejection behaviour in a heavily parasitized Great Reed Warbler Acrocephalus arundinaceus population. Ibis, 144, 614–22. Moyle, R. G., Biun, A., Butit, B., and Sumpongol, D. (2001). Brood hosts of Oriental Cuckoo Cuculus saturatus in Sabah, Malaysia. Bulletin of the British Ornithologists’ Club, 121, 107–110. Mueller, H. C. (1990). The evolution of reversed sexual dimorphism in size in monogamous species of birds. Biological Reviews, 65, 553–85. Mukherjee, A. K. (1995). Birds of arid and semi-arid tracts. Zoological Society of India Occasional Paper, Calcutta, 142.

Mundy, P. J. (1973). Vocal mimicry of their hosts by nestlings of the Great Spotted Cuckoo and Striped Crested Cuckoo. Ibis, 115, 602–4. Mundy, P. J. and Cook, A. W. (1974). Birds of Sokoto. Part 3, breeding data. Bulletin of the Nigerian Ornithologists’ Society, 10, 1–28. Mundy, P. J. and Cook, A. W. (1977). Observations on the breeding of the Pied Crow and Great Spotted Cuckoo in northern Nigeria. Ostrich, 48, 72–84. Murton, R. K. and Westwood, N. J. (1977). Avian breeding cycles. Clarendon Press, Oxford. Muth, D. P. (1991). The spring season. Central southern region. American Birds, 45, 455–60, 513. Nakamura, H. (1990). Brood parasitism by the Cuckoo Cuculus canorus in Japan and the start of new parasitism on the Azure-winged Magpie Cyanopica cyanea. Japanese Journal of Ornithology, 39, 1–18. Nakamura, H. and Miyazawa, Y. (1997). Movements, space use and social organization of radio-tracked Common Cuckoos during the breeding season in Japan. Japanese Journal of Ornithology, 46, 23–54. Nakamura, H., Kubota, S., and Suzuki, R. (1998). Coevolution between the Common Cuckoo and its major hosts in Japan. In Parasitic birds and their hosts, (eds. S. I. Rothstein and S. K. Robinson), pp. 94–112. Oxford University Press, Oxford. Narosky, T. and Yzurieta, D. (1987). Guía para la identificación de las aves de Argentina y Uruguay . Asociacion Ornitológia del Plata, Buenos Aires. Narosky, T. and Di Giacomo, A. G. (1993). Las aves de la Provincia de Buenos Aires: distribución y estatus . Asociacion Ornitólogica del Plata, Vázquez Mazzini Ed. y L. O. L. A., Buenos Aires. Nash, A. D. and Nash, S. V. (1985). Breeding notes on some Padang-Sugihan birds. Kukila, 2, 59–63. Nash, S. V. and Nash, A. D. (1988). An annotated checklist of the birds of Tanjung Puting National Park, central Kalimantan. Kukila, 3, 93–116.

Mulder, R. A. and Langmore, N. E. (1993). Dominant males punish helpers for temporary defection in Superb Fairy-wrens. Animal Behaviour, 45, 830–3.

Natarajan, V. (1993). Food and feeding habits of the southern Crow-pheasant Centropus sinensis parroti Stresemann (Aves: Cuculidae) at Pt. Calimere, Tamil Nadu. Journal of the Bombay Natural History Society, 90, 11–6.

Muller, K. A. (1971). Physical and behavioral development of a Roadrunner raised at the National Zoological Park. Wilson Bulletin, 83, 186–93.

Natarajan, V. (1997). Breeding biology of the Southern Crow-Pheasant Centropus sinensis parroti Stresemann (Aves: Cuculidae) at Point Calimere, Tamil Nadu.

Bibliography 575 Journal of the Bombay Natural History Society, 94, 56–63.

Neunteufel, A. (1951). Observaciones sobre el Dromococcyx pavoninus Pelzeln y el parasitismo de los Cuculidos. El Hornero, 9, 288–290.

Nathan, A., Legge, S., and Cockburn, A. (2001). Nestling aggression in broods of a siblicidal kingfisher, the Laughing Kookaburra. Behavioral Ecology, 12, 716–25.

Newman, K. B. (1972). A Red-chested Cuckoo grows up. Animals, 14, 448–50.

National Geographic Society. (1987). Field guide to the birds of North America. (2nd edn.) National Geographic Society, Washington, D.C.

Newman, K. and Steyn, P. (1970). [Photographs of young Red-chested Cuckoo Cuculus solitarius]. Bokmakierie, 22, 18–9.

Naumburg, E. M. B. (1930). The birds of Matto Grosso, Brazil. Bulletin of the American Museum of Natural History, 60, 1–432.

Newton, A. (1896). A dictionary of birds. Adam and Charles Black, London.

de Naurois, R. (1969). Peuplements et cycles de reproduction des oiseaux de la côte occidentale d’Afrique (du Cap Barbas, Sahara Espagnol, à la Frontière de la République de Guinée). Mémoires du Muséum National d’Histoire Naturelle, Nouvelle Série, Série A, Zoologie , 56. de Naurois, R. (1979). The Emerald Cuckoo of São Tomé and Príncipe Islands (Gulf of Guinea). Ostrich, 50, 88–93. de Naurois, R. (1994). Les oiseaux des iles du Golfo de Guinée (São Tomé, Prince et Annobon) . Instituto de Investigação Científica Tropical, Lisboa. Naylor, G. J. P. and Brown, W. M. (1998). Amphioxus mitochondrial DNA, chordate phylogeny, and the limits of inference based on comparisons of sequences. Systematic Biology, 47, 61–76. Nealen, P. M. and Ricklefs, R. E. (2001). Early diversification of the avian brain: body relationship. Journal of Zoology, 253, 391–404. Nehrkorn. A. (1894). Zur Avifauna Batjan’s. Journal für Ornithologie, 42, 156–61. Netschajew, V. A. (1977). Zur Biologie des Kleines Kuckucks im Ussurigebiet. Falke, 24, 366–71.

Newton, I. P., Adams, N. J., Brown, C. R., Enticott, J. W., and Fugler, S. R. (1983). Nonmarine vagrant birds at the Prince Edward Islands, June 1981–May 1983. Cormorant, 11, 35–8. Nicholls, J. A., Double, M. D., Rowell, D. M., and Magrath, R. D. (2000). The evolution of cooperative and pair breeding in thornbills Acanthiza (Pardalotidae). Journal of Avian Biology, 31, 165–76. Nicholls, P. (1998). Observations of adult Pallid Cuckoo feeding juveniles of the same species. Australian Bird Watcher, 17, 305–7. Nicholson, C. P. (1997). Atlas of the breeding birds of Tennessee. University of Tennessee Press, Knoxville. Nickell, W. P. (1954a). Yellow-billed Cuckoo’s egg in Mourning Dove’s nest. Wilson Bulletin, 66, 137. Nickell, W. P. (1954b). Red-wings hatch and raise a Yellow-billed Cuckoo. Wilson Bulletin, 66, 137–8. Nielsen, L. (1996). Birds of Queensland’s wet tropics and Great Barrier Reef. Gerard Industries, Bowden, South Australia. Nikolaus, G. (1987). Distribution atlas of Sudan’s birds with notes on habitat and status. Bonner Zoologisches Monographien 25.

Neufeldt, I. (1966). Life history of the Indian Cuckoo, Cuculus micropterus micropterus Gould, in the Soviet Union. Journal of the Bombay Natural History Society, 63, 399–419.

Nishida, T. and Uehara, S. (1983). Natural diet of chimpanzees (Pan troglodytes schweinfurthii): long-term record from the Mahale Mountains, Tanzania. African Study Monographs, 3, 109–30.

Neufeldt, I. A. (1968). [The eggs of the Little Cuckoo in a nest of the Bush Warbler]. Ornitologiya, 9, 365–7. [in Russian, with photograph]

Nitzsch, C. L. (1867). Nitzsch’s Pterylography. Translated from the German, edited by P. L. Sclater. Ray Society, London.

Neufeldt, I. (1971). Der Kurzflügelsänger Horeites diphone (Kittlitz). Falke, 18, 364–75.

Niven, C. K. and Niven, P. N. F. (1966). The birds of Amanzi, Uitenhage District, C. P. South African Avifauna Series, 34.

Neufeldt, I. A. (1972). [Age and sexual dimorphism in the colour of Hierococcyx fugax hyperythrus]. Ornitologiya, 10, 97–110. (in Russian)

Nolan, V. (1963). Reproductive success of birds in a deciduous scrub habitat. Ecology, 44, 305–13.

576 Bibliography Nolan, V. (1975). External differences between newly hatched cuckoos (Coccyzus americanus and C. erythropthalmus). Condor, 77, 341. Nolan, V. and Thompson, C. F. (1975). The occurrence and significance of anomalous reproductive activities in two North American non-parasitic cuckoos. Ibis, 117, 496–503. North, A. J. (1899). Descriptive catalogue of the nests & eggs of birds found breeding in Australia and Tasmania. Australian Museum Catalogue, 12, Sydney. North, M. E. W. (1958). Voices of African birds. Laboratory of Ornithology, Cornell University, Ithaca, New York.

Oberholser, H. C. (1917). The birds of the Anamba Islands. United States National Museum Bulletin, 98, 1–75. Oberholser, H. C. (1974). Bird life of Texas, vol. 1. University of Texas Press, Austin. O’Donel, H. V. (1936). The Indian Cuckoo (Cuculus m. micropterus Gould). Journal of the Bombay Natural History Society, 39, 175. Ogawa, M. (1905). Notes on Mr. Alan Owston’s collections of birds from the islands lying between Kiushu and Formosa. Annotationes Zoologicae Japonenses, 5, 175–232, with 3 plates.

Noske, R. (1994). Shining Bronze-Cuckoo and Channel-billed Cuckoo: first records for Timor. Kukila, 7, 68–9.

Ogilvie-Grant, W. R. (1915). Report on the birds collected by the British Ornithologists’ Union Expedition and the Wollaston Expedition in Dutch New Guinea. Ibis, Jubilee Supplement, 2.

Novaes, F. C. and Cunha Lima, M. de F. (1998). Aves da Grande Belém, Municípos de Belém e Ananindeua, Pará. Coleção Emilie Snethlage, Museu Paraense Emílio Goeldi, Belém.

Ogilvie-Grant, W. R. and Forbes, H. O. (1903). Birds of Sokotra and ’Abd al-Kuri. In The natural history of Sokotra and ’Abd al-Kuri, (ed. H. O. Forbes), 19–72. Porter, London.

Nur, U. (1970). Evolutionary rates of models and mimics in Batesian mimicry. American Naturalist, 104, 477–86.

Ohmart, R. D. (1972). Physiological and ecological observations concerning the salt-secreting nasal glands of the Roadrunner. Comparative Biochemistry and Physiology, A, 43, 311–6.

Oates, E. W. (1877). Notes on the nidification of some Burmese birds. Stray Feathers, 5, 141–70. Oates, E. W. (1887). A list of the birds of Pegu. Stray Feathers, 10, 175–248. Oates, E. W. (1903). Catalogue of the collection of birds’ eggs in the British Museum (Natural History), vol. 3. British Museum (Natural History), London. Oates, E. W. and Blanford, W. T. (1895). The fauna of British India, including Ceylon and Burma. Birds, vol. 3. Taylor and Francis, London. Oates, J. F. (1999). Myth and reality in the rain forest, how conservation strategies are failing in West Africa. University of California Press, Berkeley. Oatley, T. B. (1970). Robin hosts of the Red-chested Cuckoo in Natal. Ostrich, 41, 232–6. Oatley, T. B. (1980). Eggs of two cuckoo genera in one nest, and a new host for Emerald Cuckoo. Ostrich, 51, 126–7. Oberholser, H. C. (1912). Descriptions of one hundred and four new species and subspecies of birds from the Barussan Islands and Sumatra. Smithsonian Miscellaneous Collections, 60 (7), 1–22.

Ohmart, R. D. (1973). Observations on the breeding adaptations of the Roadrunner. Condor, 75, 140–9. Ohmart, R. D. (1989). A timid desert creature that appears to be half bird, half reptile. Natural History, 98 (9), 34–40. Ohmart, R. D. and Lasiewski, R. C. (1971). Roadrunners: energy conservation by hypothermia and absorption of sunlight. Science, 172, 67–9. Ohmart, R. D., Chapman, T. E., and McFarland, L. Z. (1970a). Water turnover in Roadrunners under different environmental conditions. Auk, 87, 787–93. Ohmart, R. D., McFarland, L. Z., and Morgan, J. T. (1970b). Urographic evidence that urine enters the rectum and ceca of the Roadrunner. Comparative Biochemistry and Physiology A, 35, 487–9. Øien, I. J., Moksnes, A., and Røskaft, E. (1995). Evolution of variation in egg color and marking pattern in European passerines: adaptations in a coevolutionary arms race with the Cuckoo, Cuculus canorus. Behavioral Ecology, 6, 166–74. Øien, I. J., Honza, M., Moksnes, A., and Røskaft, E. (1996). The risk of parasitism in relation to the distance

Bibliography 577 from Reed Warbler nests to Cuckoo perches. Journal of Animal Ecology, 65, 147–53. Øien, I. J., Moksnes, A., Røskaft, E., and Honza, M. (1998). Costs of Cuckoo Cuculus canorus parasitism to Reed Warblers Acrocephalus scirpaceus. Journal of Avian Biology, 29, 205–15. Oliver, W. R. B. (1955). New Zealand birds. A. H. and A. W. Reed, Wellington. Olson, S. L. (1975). Paleornithology of St. Helena Island, South Atlantic Ocean. Smithsonian Contributions to Paleobiology, 23, 1– 49.

Omland, K. E., Lanyon, S. M, and Fritz, S. J. (1999). A molecular phylogeny of the New World orioles (Icterus): the importance of dense taxon sampling. Molecular Phylogenetics and Evolution, 12, 224–39. Oren, D. C. and Parker, T. A. (1997). Avifauna of the Tapajós National Park and vicinity, Amazonian Brazil. Ornithological Monographs, 48, 493–525. Oniki, Y. and Ricklefs, R. E. (1981). More growth rates of birds in the humid New World tropics. Ibis, 123, 349–54.

Olson, S. L. (1978). Greater Ani (Crotophaga major) in Mexico. Auk, 95, 766 –7.

Oniki, Y. and Willis, E. O. (1972). Studies of antfollowing birds north of the eastern Amazon. Acta Amazônica, Manaus 2, 127–51.

Olson, S. L. (1983). Evidence for a polyphyletic origin of the Piciformes. Auk, 100, 126–33.

Orenstein, R. I. (1976). Birds of the Plesyumi area, central New Britain. Condor, 78, 370–4.

Olson, S. L. (1985). The fossil record of birds. In Avian Biology, vol. 8 (eds. D. S. Farner, J. R. King, and K. C. Parkes), pp. 79–238. Academic Press, New York.

Osborne, T. O. (1973). Additional notes on the birds of the Kafue Flats. Puku, 7, 163–6.

Olson, S. L. (1990). Future paleontological potential of St. Helena. In St. Helena natural treasury, (eds. P. Pearce-Kelly and Q. C. B. Cronk), pp. 11–2. Proceedings of a symposium held at the Zoological Society of London 9th September 1988. Zoological Society of London, London. Olson, S. L. (1992). A new family of primitive landbirds from the Lower Eocene Green River formation of Wyoming. In Papers in Avian Paleontology, honoring Pierce Brodkorb. Natural History Museum of Los Angeles County, Science Series no. 36, 127–36. Olson, S. L. and Feduccia, J. A. (1979). An Old World occurrence of the Eocene avian family Primobucconidae. Proceedings of the Biological Society of Washington, 92, 494–7. Olson, S. L. and Hilgartner, W. B. (1982). Fossil and subfossil birds from the Bahamas. Smithsonian Contributions in Paleobiology, 48, 22–60. Olson, S. L. and James, H. F. (1982). Prodromus of the fossil avifauna of the Hawaiian islands. Smithsonian Contributions in Zoology, 365, 1–59.

Osmaston, A. E. (1912). Eggs of the Large Hawkcuckoo (Hierococcyx sparverioides). Journal of the Bombay Natural History Society, 21, 1330–1. Osmaston, B. B. (1916). Notes on cuckoos in Maymyo. Journal of the Bombay Natural History Society, 24, 359–63. Ostende, L. W. van den Hoek, Dekker, R. W. R. J., and Keijl, G. O. (1997). Type-specimens of birds in the National Museum of Natural History, Leiden. Part 1. Non-passerines. National Museum of Natural History, Leiden,Technical Bulletin, 1. Ottow, J. and Duve, G. (1965). Zur Kenntnis der Fortpflanzung von Chrysococcyx caprius und Cuculus canorus gularis in Süd-Africa. Journal für Ornithologie , 106, 431–9. Ottow, J. and Verheijen, J. A. J. (1969). Zur Lebensweise der Kuckucke von Flores. Journal für Ornithologie , 110, 27–9. Owen, R. (1861). On the nesting of some Guatemalan birds. Ibis, 3, 58–69. Oustalet, E. (1899). Les oiseaux du Cambodge, du Laos, de l’Annam et du Tonkin. Part 1. Nouvelles et Archives du Muséum de Paris, Ser. 4, 1, 221–96.

Olson, S. L. and James, H. F. (1991). Descriptions of thirty-two new species of birds from the Hawaiian Islands: Part I. Non-passeriformes. Ornithological Monographs, 45.

Pagel, T. (1992). Kuckucksvögel—selten gepflegt. Tropische Vögel , 13, 46–52.

Olsson, V. (1948). Om en gökunges (Cuculus canorus) tillväxt. Några iakttagelser och jämförelser [On the growth of a young Cuckoo). Vår Fågelvärld , 7, 49–56.

Pakenham, R. H. W. (1979). The birds of Zanzibar and Pemba. BOU Check-list No. 2, British Ornithologists’ Union, London.

578 Bibliography Palomino, J. J., Martin-Vivaldi, M., Soler, M., and Soler, J. J. (1998). Females are responsible for ejection of Cuckoo eggs in the Rufous Bush Robin. Animal Behaviour, 56, 131–6. Paludan, K. (1938). Zur Ornis des Zagrossgebietes,W.Iran. Journal für Ornithologie , 86, 562– 638. Paludan, K. (1959). On the birds of Afghanistan. The 3rd Danish Expedition to Central Asia, Zoological results 25. Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening i København , Bind 122. Pandey, S., Joshua, J., Rai, N. D., Mohan, D., Rawat, G. S., Sankar, K., Katti, M. V., Khati, D. V. S., and Johnsingh, A. J. T. (1994). Birds of Rajaji National Park, India. Forktail, 10, 105–13. Park, H.-S. and Kim, W.-B. (1995). First records of Chestnut-winged Cuckoo (Clamator coromandus), European Starling (Sturnus vulgaris) and Black Bittern (Ixobrychus flavicollis) in Korea. Korean Journal of Ornithology, 2, 75–6. Parker, A. R., McKenzie, D. R., and Large, M. C. J. (1998). Multilayer reflectors in animals using green and gold beetles as contrasting examples. Journal of Experimental Biology, 201, 1307–13.

Parkes, K. C. (1973). Annotated list of the birds of Leyte Island, Philippines. Nemouria, 11. Parkes, K. C. (1984). An apparent hybrid Black-billed ⫻ Yellow-billed Cuckoo. Wilson Bulletin, 96, 294–6. Parkes, K. C. and Niles, D. M. (1988). Notes on Philippine birds, 12. An undescribed subspecies of Centropus viridis. Bulletin of the British Ornithologists’ Club, 108, 193–4. Payne, R. B. (1965). Clutch size and numbers of eggs laid by Brown-headed Cowbirds. Condor, 67, 44–60. Payne, R. B. (1967). Interspecific communication signals in parasitic birds. American Naturalist, 101, 363–76. Payne, R. B. (1968). The birds of mopane woodland and other habitats of Hans Merensky Nature Reserve, Transvaal. South African Avifauna Series, 56. Payne, R. B. (1969a). Overlap of breeding and molting schedules in a collection of African birds. Condor, 71, 140–5. Payne, R. B. (1969b). The breeding seasons and reproductive physiology of Tricolored Blackbirds and Redwinged Blackbirds. University of California Publications in Zoology, 90, 1–137. Payne, R. B. (1969c). Nest parasitism and display of Chestnut Sparrows in a colony of Grey-capped Social Weavers. Ibis, 111, 300–7.

Parker, S. A. (1981). Prolegomenon to further studies in the Chrysococcyx ‘malayanus’ group (Aves, Cuculidae). Zoologische Verhandelingen, Rijksmuseum van Natuurlijke Historie te Leiden, 187.

Payne, R. B. (1971). Duetting and chorus singing in African birds. Ostrich, Supplement, 9, 125–45.

Parker, T. A. (1982). Observations of some unusual rainforest and marsh birds in southeastern Peru. Wilson Bulletin, 94, 477–93.

Payne, R. B. (1973a). Individual laying histories and the clutch size and numbers of eggs of parasitic cuckoos. Condor, 75, 414–38.

Parker, T. A. and Goerck, J. M. (1997). The importance of National Parks and Biological Reserves to bird conservation in the Atlantic forest region of Brazil. Ornithological Monographs, 48, 527–41.

Payne, R. B. (1973b). The breeding season of a parasitic bird, the Brown-headed Cowbird, in central California. Condor, 75, 80–99.

Parker, T. A. and Remsen, J. V. (1987). Fifty-two Amazonian bird species new to Bolivia. Bulletin of the British Ornithologists’ Club, 107, 94 –107. Parker, V. (1994). Swaziland bird atlas 1985 –1991 . Websters, Mbabane, Swaziland. Parker, V. (1999). The atlas of the birds of Sul do Save, southern Mozambique. Avian Demography Unit, Cape Town. Parkes, K. C. (1971). Taxonomic and distributional notes on Philippine birds. Nemouria, 4, 1– 64.

Payne, R. B. (1974). The evolution of clutch size and reproductive rates in parasitic cuckoos. Evolution, 28, 169–81. Payne, R. B. (1977a). Juvenile plumages of Cuculus canorus and Cuculus gularis in Africa. Bulletin of the British Ornithologists’ Club, 97, 48–53. Payne, R. B. (1977b). The ecology of brood parasitism in birds. Annual Review of Ecology and Systematics 8, 1–28. Payne, R. B. (1977c). Clutch size, egg size, and the consequences of single vs. multiple parasitism in parasitic finches. Ecology, 58, 500–13.

Bibliography 579 Payne, R. B. (1983). Bird songs, sexual selection, and female mating strategies. In Social behavior of female vertebrates (ed. Wasser, S. K.), pp. 55–90. Academic Press, New York.

Payne, R. B. and Sorenson, M. D. (2003). Museum collections as sources of genetic data. Bonnischer Zoological Beitrage, 51, 97–104.

Payne, R. B. (1984). Sexual selection, lek and arena behavior, and sexual size dimorphism in birds. Ornithological Monographs, 33.

Payne, R. B., Payne, L. L., and Rowley, I. (1985). Splendid Wren Malurus splendens response to cuckoos: an experimental test of helping and social organization in a communal bird. Behaviour, 94, 108–27.

Payne, R. B. (1986). Bird song and avian systematics. Current Ornithology, 3, 87–125. Payne, R. B. (1989). Egg size of African honeyguides (Indicatoridae): specialization for brood parasitism? Tauraco, 1, 201–10. Payne, R. B. (1990). Natal dispersal, area effects, and effective population size. Journal of Field Ornithology, 61, 396 – 403. Payne, R. B. (1997a). Avian brood parasitism. In Hostparasite evolution: general principles and avian models, (eds. D. H. Clayton and J. Moore), pp. 338– 69. Oxford University Press, Oxford. Payne, R. B. (1997b). Family Cuculidae (Cuckoos). In Handbook of the birds of the world, volume 4, (eds. J. del Hoyo, A. Elliott, and J. Sargatal), pp. 508– 607. BirdLife/Lynx Edicions, Barcelona. Payne, R. B. (1998). Brood parasitism in birds: strangers in the nest. BioScience, 48, 377–86. Payne, R. B. and Payne, K. (1967). Cuckoo hosts in southern Africa. Ostrich, 38, 135–43.

Payne, R. B., Payne, L. L., and Rowley, I. (1988). Kinship and nest defence in cooperative birds: Splendid Fairywrens Malurus splendens. Animal Behaviour, 36, 939–41. Payne, R. B., Payne, L. L., and Woods, J. L. (1998). Song learning in brood parasitic indigobirds Vidua chalybeata: song mimicry of the host species. Animal Behaviour, 55, 1537–53. Payne, R. B., Payne, L. L., Woods, J. L., and Sorenson, M. D. (2000). Imprinting and the origin of parasite-host species associations in brood parasitic indigobirds Vidua chalybeata. Animal Behaviour, 59, 69–81. Payne, R. B., Woods, J. L., and Payne, L. L. (2001). Parental care in estrildid finches: experimental tests of a colonization model of Vidua brood parasitism. Animal Behaviour, 62, 473–83. Payne, R. B., Hustler, K., Stjernstedt, R., Sefc, K. M., and Sorenson, M. D. (2002). Behavioural and genetic evidence of a recent population switch to a novel host species in brood parasitic indigobirds Vidua chalybeata. Ibis, 144, 373–83.

Payne, R. B. and Payne, L. L. (1998a). Nestling eviction and vocal begging behaviors in Australian glossy cuckoos Chrysococcyx basalis and C. lucidus. In Parasitic birds and their hosts, (eds. S. I. Rothstein and S. K. Robinson), pp. 152–69. Oxford University Press, Oxford.

Paynter, R. A., Jr. (1955). The ornithogeography of the Yucatán peninsula. Peabody Museum of Natural History, Yale University, Bulletin 9.

Payne, R. B. and Payne, L. L. (1998b). Brood parasitism by Brown-headed Cowbirds: risks and costs on reproductive success and survival in Indigo Buntings. Behavioral Ecology, 9, 64–73.

Peck, G. K. and James, R. D. (1983). Breeding birds of Ontario, nidiology and distribution. vol. 1: nonpasserines. Life Sciences Miscellaneous Publications, Royal Ontario Museum.

Payne, R. B. and Payne, L. L. (2002). Begging for parental care from another species: specialization and generalization in brood-parasitic finches. In The evolution of begging: competition, cooperation and communication, (eds. J. Wright and M. L. Leonard), pp. 429–50. Kluwer Academic, Doordrecht.

Pelzeln, A. von. (1871). Zur Ornithologie Brasiliens. Zoologischen Cabinete, Wien.

Payne, R. B. and Risley, C. J. (1976). Systematics and evolutionary relationships among the herons. University of Michigan Miscellaneous Publications of the Museum of Zoology, 150.

Pearson, R. (1972). The avian brain. Academic Press, New York.

Pemberton, J. R. (1916). Variation of the broken-wing stunt by a Roadrunner. Condor, 18, 203. Pemberton, J. R. (1925). Parasitism in the Road-runner. Condor, 27, 35. de la Peña, M. R. (1983). Nota sobre observaciones de nidification de aves en la Provincia de Santa Fe. Hornero, 12, 129–31.

580 Bibliography de la Peña, M. R. (1986). Guia de aves Argentinas, tomo IV, Columbiformes a Piciformes. Luis Huber, Dibujos, Argentina. de la Peña, M. R. (1993). Parasitismo del crespin, Tapera naevia chochi (Vieillot, 1817) en la Provincia de Santa Fe, Republica Argentina. Notulas Faunisticas, 37, 1– 4. de la Peña, M. R. (1995). Ciclo reproductivo de las aves Argentinas. L. O. L. A., Buenos Aires. de la Peña, M. R. and Rumboll, M. (1998). Collins illustrated checklist: birds of southern South America and Antarctica. HarperCollins, London. Penard, F. P. and Penard, A. P. (1910). De Vogels van Guyana (Suriname, Cayenne en Demerara), vol. 2. F. P. Penard, Paramaribo. Pérez de Val, J. (1996). Las aves de Bioko Guinea Equatorial, guía de campo. Edilesa, León. Pérez de Val, J., Castroviejo, J., and Purroy, F. F. (1997). Species rejected from and added to the avifauna of Bioko island (Equatorial Guinea). Malimbus, 19, 19–31. Perrin de Brichambaut, J. (1993). Mimetisme des oeufs de coucou gris. Alauda, 61, 161–72. Perrins, C. (1967). The short apparent incubation period in the Cuckoo. British Birds, 60, 50 –1. Peterjohn, B. G., Sauer, J. R., and Robbins, C. S. (1995). Population trends from the North American Breeding Bird Survey. In Ecology and management of Neotropical migratory birds, (eds. T. E. Martin and D. M. Finch), pp. 3 –39. Oxford University Press, Oxford. Peters, J. L. (1940). Check-list of birds of the world. vol. 4. Museum of Comparative Zoology, Cambridge, Massachusetts. Peterson, R. T. (1980). A field guide to the birds. (4th edn.) Houghton Mifflin, Boston. Pettet, A. (1975). Breeding behaviour of Clamator glandarius at Ibadan in southern Nigeria. Bulletin of the Nigerian Ornithologists’ Society, 11, 34 – 40. Phelps, W. H. and Phelps, W. H., Jr. (1958). Lista de las aves de Venezuela con su distribucion. Tomo II, Parte 1, No Passeriformes. Boletin de la Sociedad Venezolana de Ciencias Naturales, 19 (90), 1–317. Phillipi, T. and Seger, J. (1989). Hedging one’s evolutionary bets, revisited. Trends in Ecology and Evolution, 4, 41–4. Phillips, N. J. (1984). Migrant species new to Seychelles. Bulletin of the British Ornithologists’ Club, 104, 9–10.

Phillips, W. W. A. (1948). Cuckoo problems of Ceylon. Spolia Zeylanica, 25, 45–60. Phillips, W. W. A. (1949). A new race of the Common Hawk Cuckoo from Ceylon. Bulletin of the British Ornithologists’ Club, 69, 56–7. Phillipps, A. (1970). Some important nesting notes from Sabah. Sabah Society Journal, 5, 141–4. Picman, I. (1989). Mechanism of increased puncture resistance of eggs of Brown-headed Cowbirds. Auk, 106, 577–83. Picman, J. and Pribil, S. (1997). Is greater eggshell density an alternative mechanism by which parasitic cuckoos increase the strength of their eggs? Journal für Ornithologie, 138, 531–41. Piechocki, R. (1971). Die Mauser der Kuckucke Saurothera merlini und Crotophaga ani. Ergebnisse der 1. Kubanisch Deutschen “Alexander-von Humboldt” Expedition 1967/68. Zoologische Jahrbuch (Systematik), 98, 1–10. Pigram, C. J. and Davies, H. L. (1987). Terranes and the accretion history of the New Guinea orogen. BMR Journal of Australian Geology and Geophysics, 10, 193–211. Pinto, A. A. da Rosa. (1959). Um esbocada avifauna sedentario da regiao da Gorongoza, Moçambique. Ostrich Suppl., 3, 98–125. Pinto, A. A. da Rosa. (1983). Ornitologia de Angola, vol. 1. Instituto de Investigação Científica Tropical, Lisboa. Pinto, O. M. de O. (1938). Catalogo das Aves do Brasil. Revista do Museu Paulista, Universidade de Saõ Paulo , 22, Saõ Paulo. Pinto, O. M. de O. (1947). Contribuição à ornitologia do Baixo Amazonas. Estudo críto de uma coleção de aves do Estado do Pará. Arquivos de Zoologia do Estado de São Paulo , 5, 311–482. Pinto, O. M. de O. (1952). Súmula histórica e sistemática da ornitologia de Minas Gerais. Arquivos de Zoologia do Estado de São Paulo , 8, 1–51. Pinto, O. M. de O. (1953). Sobre a coleção Carlos Estevão de Peles, ninhos e ovos de aves de Belém (Pará). Papéis Avulsos do Departamento de Zoologia, Secretaria da Agricultura, São Paulo, Brasil , 11, 111–222. Pinto, O. M. de O. (1962). Notas sobre a variaçao geográfica nas populações Brasileiras de Neomorphus geoffroyi, com a descrição de uma subespécie nova. Papéis Avulsos do Departamento de Zoologia, Secretaria da Agricultura, São Paulo, Brasil , 15, 299–300.

Bibliography 581 Pinto, O. M. de O. (1978). Novo catálogo das aves do Brasil. Empresa Gráfica da Revista dos Tribunais, Saõ Paulo.

Posso, S. R. (2003). Osteologia craniana de Cuculiformes (Aves) com implicaçiões evolutivas. PhD Dissertation, Universidade de São Paulo.

Pistorius, A. (1985). Yellow-billed Cuckoo. In The atlas of breeding birds in Vermont, (eds. S. B. Laughlin and D. P. Kibbe), pp. 128–9. Vermont Institute of Natural Science, University Press of New England, Hanover, New Hampshire.

Posso, S. R. and Donatelli, R. J. (2001). Cranial osteology and systematic implications in Crotophaginae (Aves, Cuculidae). Journal of Zoology and Systematic Evolutionary Research, 39, 247–56.

Pitman, C. R. S. (1928). The breeding of Centropus superciliosus loandae—the central African White-browed Coucal in Uganda. Oologists’ Record, 8, 41–3. Pitman, C. R. S. (1957). On the egg of the African Cuckoo Cuculus canorus gularis Stephens. Bulletin of the British Ornithologists’ Club, 77, 138–9. Pizzey, G. (1980). A field guide to the birds of Australia. Princeton University Press, Princeton. Pizzey, G. and Knight, F. (1998). Collins field guide to the birds of Australia. Collins, London. Pleasants, J. M. and Pleasants, B. Y. (1989). Reversed size dimorphism in raptors: can we tell which sex changed size? Oikos, 56, 287–8. Plowes, D. H. C. (1948). Young African Cuckoo in drongo nest. Ostrich, 19, 99–100. Poiani, A. (1998). Did avian heterospecific brood parasites evolve from a cooperatively breeding ancestor? Ethology, Ecology and Evolution, 10, 253–75. Poiani, A. and Elgar, M. A. (1994). Cooperative breeding in the Australian avifauna and brood parasitism by cuckoos (Cuculidae). Animal Behaviour, 47, 697–706. Polhemus, D. A. and Polhemus, J. T. (1998). Assembling New Guinea: 40 million years of island arc accretion as indicated by the distributions of aquatic Heteroptera (Insecta). In Biogeography and geological evolution of SE Asia, (eds. R. Hall and J. D. Holloway), pp. 327– 40. Backhuys, Leiden, The Netherlands. Portenko, L. (1931). Einige neue Unterarten paläarktischer Vögel. Mitteilungen aus der Zoologischen Museum in Berlin, 17, 415–23. Porter, R. F., Willis, I., Christensen, S., and Nielsen, B. P. (1976). Flight identification of European raptors, 2nd ed. T. and A. B. Poyser, Birkhamstead, UK. Porter, R. F., Martins, R. P., Shaw, K. D., and Sørensen, U. (1996). The status of non-passerines in southern Yemen and the records of the OSME survey in spring 1993. Sandgrouse, 17, 22–53.

Potter, R. F. (1980). Notes on nesting Yellow-billed Cuckoos. Journal of Field Ornithology, 51, 17–29. Potter, R. F. (1981). Effects of cool weather on nesting behavior and development in the Yellow-billed Cuckoo. Chat, 45, 15–6. Poulin, B., Lefebvre, G., and McNeil, R. (1992). Tropical avian phenology in relation to abundance and exploitation of food resources. Ecology, 73, 2295–309. Poulsen, M. K. (1995). The threatened and near threatened birds of Luzon, Philippines, and the role of the Sierra Madre mountains in their conservation. Bird Conservation International, 5, 79–115. Poulsen, M. K. and Lambert, F. R. (2000). Altitudinal distribution and habitat preferences of forest birds on Halmahera and Buru, Indonesia: implications for conservation of Moluccan avifaunas. Ibis, 142, 566–86. Pratt, E. (1972). Some answers about Coucal nesting. Sunbird, 3, 19–25. Pratt, H. D., Bruner, P. L., and Berrett, D. G. (1977). Ornithological observations on Yap, Western Caroline Islands. Micronesica, 13, 59–6. Prawiradilaga, D. M., Marakarmah, A., and Wijamukti, S. (2003). A photographic guide to the birds of Javan montane forest: Gunung Halimun National Park. Biodiversity Conservation Project, LIPI, Bogor. Preble, N. A. (1957). Nesting habits of the Yellow-billed Cuckoo. American Midland Naturalist, 57, 474– 82. Price, J., Droge, S., and Price, A. (1995). The summer atlas of North American breeding birds. Academic Press, London. Price, T. D. (1979). The seasonality and occurrence of birds in the Eastern Ghats of Andhra Pradesh. Journal of the Bombay Natural History Society, 76, 379–422. Price, T. D. and Jamdar, N. (1990). The breeding birds of Overa Wildlife Sanctuary, Kashmir. Journal of the Bombay Natural History Society, 87, 1–15. Price, T. D. and Jamdar, N. (1991). Breeding of eight sympatric species of Phylloscopus warblers in Kashmir. Journal of the Bombay Natural History Society, 88, 242–55.

582 Bibliography Promptov, A. N. and Loukina, E. W. (1940). Sur les rapports biologiques réciproques du coucou (Cuculus canorus L.) avec certain espèces d’oiseaux qui l’élèvent ses petits. Bulletin de Société Naturelle Moscou Section de Biologie, 49, 82–96. (in Russian, French résumé). Pruett, C. L., Gibson, D. D., and Winker, K. (2001). Molecular “cuckoo clock” suggests listing of western Yellow-billed Cuckoos may be warranted. Wilson Bulletin, 113, 228–31. Prum, R. O. and R. Torres. (2003). Structural colouration of avian skin: convergent evolution of coherently scattering dermal collagen arrays. Journal of Experimental Biology, 206, 2409–29. Prum, R. O., Torres, R. H., Williamson, S., and Dyck, J. (1998). Coherent light scattering by blue feather barbs. Nature, 396, 28–9. Prum, R. O., Torres, R., Williamson, S., and Dyck, J. (1999). Two-dimensional Fourier analysis of the spongy medullary keratin of structurally coloured feather barbs. Proceedings of the Royal Society of London, Series B, 266, 13 –22.

Quickelberge, C. D. (1989). Birds of the Transkei. Durban Natural History Museum, Durban. Quinn, J. S. and Startek-Foote, J. M. (2000). Smoothbilled Ani Crotophaga ani. In The Birds of North America no. 539, (eds. A. Poole and F. Gill). The Birds of North America, Philadelphia. Quinn, J. S., Macedo, R., and White, B. N. (1994). Genetic relatedness of communally breeding Guira Cuckoos. Animal Behaviour, 47, 515–29. Rabor, D. S. (1977). Philippine birds and mammals. Science Education Center, University of the Philippines Press, Quezon City. Raethel, H.-S. (1992). Der Schuppenhelmkuckuck Phoenicophaeus cumingi. Gefiederte Welt, 16, 11–2. Raffaele, H. A. (1989). A guide to the birds of Puerto Rico and the Virgin Islands. Princeton University Press, Princeton. Raffaele, H. A., Wiley, J., Garrido, O., Keith, A., and Raffaele, J. (1998). Birds of the West Indies. Christopher Helm, London.

Pryce, E. (1989). A Black Cuckoo raised by Swamp Boubous. Babbler, Gaborone, 18, 38.

Raffles, T. S. (1817). The history of Java. Black, Parbury and Allen, London.

Prys-Jones, R. and Diamond, A. W. (1984). Ecology of the land birds on the granitic and coralline islands of the Seychelles, with particular reference to Cousin Island and Aldabra Atoll. In Biography and ecology of the Seychelles Islands, (ed. D. R. Stoddart). Monographiae Biologicae, 55, 529–58. Dr. W. Junk, The Hague.

Raffles, T. S. (1822). Second part of the descriptive catalogue of a zoological collection made in the island of Sumatra and its vicinity. Transactions of the Linnean Society of London, 13, 277–340.

Punnett, R. C. (1933). Inheritance of egg colour in the ‘parasitic’ cuckoo. Nature, 132, 892–3. Pycraft, W. P. (1901). Some points in the morphology of the palate of the Neognathae. Journal of the Linnean Society, Zoology, 28, 343 –57. Pycraft, W. P. (1903). Contributions to the osteology of birds, part vi. Cuculiformes. Proceedings of the Zoological Society of London 1903, vol. 1, 258 –91. Pyle, P. W. (1997). Identification guide to North American birds, part 1. Slate Creek Press, Bolinas, California. Quay, W. B. (1967). Comparative survey of the anal glands of birds. Auk, 84, 379–89. Queller, D. C., Strassmann, J. E., and Hughes, C. R. (1988). Genetic relatedness in colonies of tropical wasps with multiple queens. Science, 242, 1155–7.

Raffles, S. (1830). Memoir of the life and public services of Sir Stamford Raffles. Catalogue of Zoological Specimens, Aves, 648–87. Rahn, H. and Paganelli, C. V. (1989). Shell mass, thickness and density of avian eggs derived from the tables of Schönwetter. Journal für Ornithologie , 130, 59–68. Raikow, R. J. (1985). Locomotor system. In Form and function in birds, vol. 3, (eds. A. S. King and J. McLelland), pp. 57–147. Academic Press, New York. Raikow, R. J. and Cracraft, J. (1983). Monophyly of the Piciformes: a reply to Olson. Auk, 100, 134–8. Rajasingh, S. G. and Rajasingh, I. V. (1970). Birds and mammals eating the fruits of yellow oleander (Thevetia peruviana). Journal of the Bombay Natural History Society, 67, 572–3. Ralph, C. P. (1975). Life style of Coccyzus pumilus, a tropical cuckoo. Condor, 77, 60–72.

Bibliography 583 Rand, A. L. (1933). Testicular asymmetry in the Madagascar Coucal. Auk, 50, 219–20.

basin: regulation by long-term sub-Andean tectonics. Science, 238, 1398–401.

Rand, A. L. (1936). The distribution and habits of Madagascar birds. Bulletin of the American Museum of Natural History, 72, 143 – 499.

Rasmussen, P. C. and Prys-Jones, R. P. (2003). History vs mystery: the reliability of museum specimen data. Bulletin of the British Ornithologists’ Club, 123A, 66–94.

Rand, A. L. (1941a). Results of the Archbold Expeditions. no. 32. New and interesting birds from New Guinea. American Museum Novitates, 1192. Rand, A. L. (1941b). Courtship of the Roadrunner. Auk, 58, 57–9. Rand, A. L. (1942a). Results of the Archbold Expeditions. no. 42. Birds of the 1936 –1937 New Guinea Expedition. Bulletin of the American Museum of Natural History, 79, 289–366. Rand, A. L. (1942b). Results of the Archbold Expeditions. no. 43. Birds of the 1938–1939 New Guinea Expedition. Bulletin of the American Museum of Natural History, 79, 425–515. Rand, A. L. (1951). Birds of Negros Island. Fieldiana: Zoology, 31, 571–96. Rand, A. L. (1953). Factors affecting feeding rates of anis. Auk, 70, 26–30. Rand, A. L. and Fleming, R. L. (1957). Birds from Nepal. Fieldiana: Zoology, 41, 1–218. Rand, A. L. and Gilliard, E. T. (1967). Handbook of New Guinea birds. Weidenfeld and Nicholson, London. Rand, A. L. and Rabor, D. S. (1960). Birds of the Philippine Islands: Siquijor, Mount Malindang, Bohol, and Samar. Fieldiana: Zoology, 35, 222–441.

Raxworthy, C. J., Forstner, M. R. J., and Nussbaum, R. A. (2002). Chameleon radiation by oceanic dispersal. Nature, 415, 784–7. Ray-Chaudhuri, R. (1967). Mitotic and meiotic chromosomes of the Koel Eudynamys scolopacea scolopacea. Nucleus, 10, 179–89. Ray-Chaudhuri, R. (1973). Cytotaxonomy and chromosome evolution in birds. In Cytotaxonomy and vertebrate evolution, (eds. A. B. Chiarelli and E. Capannan), pp.425–83. Academic Press, London. Read, J. (1995). First South Australian record of the Oriental Cuculus saturatus. South Australian Ornithologist, 32, 62–3. Reboreda, J. C., Clayton, N. S., and Kacelnik, A. (1996). Species and sex differences in hippocampus size in parasitic and non-parasitic cowbirds. NeuroReport, 7, 505–8. Recher, H. F. and Holmes, R. T. (1985). Foraging ecology and seasonal patterns of abundance in a forest avifauna. In Birds of eucalypt forests and woodlands: ecology, conservation, management, (eds. A. Keast, H. F. Recher, H. A. Ford, and D. Saunders), pp. 79–96. RAOU, Surrey Beatty, Sydney.

Rand, A. L. and Rabor, D. S. (1967). New birds from Luzon, Philippine Islands. Fieldiana: Zoology, 51, 85–9.

Redondo,T. (1993). Exploitation of host mechanisms for parental care by avian brood parasites. Etologia, 3, 235–97.

Rand, A. L., Friedmann, H., and Traylor, M. A. (1959). Birds from Gabon and Moyen Congo. Fieldiana: Zoology, 41, 221– 411.

Redondo, T. and Arias de Reyna, L. (1988). Vocal mimicry of hosts by Great Spotted Cuckoo Clamator glandarius: further evidence. Ibis, 130, 540–4.

Randrianary, V., Rifflet, S., and Roché, J. C. (1997). Madagascar soundscapes. (sound disk, CD). Sitelle, Le Verdier, France.

Redondo, T. and Arias de Reyna, L. (1989). High breeding success in experimentally parasitized broods of Azure-winged Magpies (Cyanopica cyanea). Gerfaut, 79, 149–52.

Rappole, J. H., Morton, E. S., Lovejoy, T. E., and Ruos, J. L. (1993). Aves migratorias nearcticas en los Neotropicos. Smithsonian Institution, Front Royal, Virginia. Räsänen, M. E., Salo, J. S., and Kalliola, R. J. (1987). Fluvial perturbance in the western Amazon

Reed, R. A. (1953). Some observations on the Didric Cuckoo (Chrysococcyx caprius). Ostrich, 24, 138–40. Reed, R. A. (1968). Studies of the Diederik Cuckoo Chrysococcyx caprius in the Transvaal. Ibis, 110, 321–31.

584 Bibliography Reed, R. A. (1969). Notes on the Redchested Cuckoo in the Transvaal. Ostrich, 40, 1–4.

Ricklefs, R. E. (1969). An analysis of nesting mortality in birds. Smithsonian Contributions in Zoology, 9, 1–48.

Reeve, H. K., Emlen, S. T., and Keller, L. (1998). Reproductive sharing in animal societies: reproductive incentives or incomplete control by dominant breeders? Behavioral Ecology, 9, 267–78.

Ricklefs, R. E. (1973). Patterns of growth in birds. II. Growth rate and mode of development. Ibis, 115, 177–201.

Reichholf, J. (1974). Biotopwahl und Schwarmgrößen bei den neotropischen Kuckucksvögeln Crotophaga ani L. und Guira guira (Gmel.). Bonner Zoologische Beiträge , 25, 118–22.

Riddell, I. C. (1990). Young Thick-billed Cuckoo at Mana Pools. Honeyguide, 36, 142–3. Riddell, I. C. (1994). Red-necked Falcon and Thick-billed Cuckoo in the Zambezi Valley. Honeyguide, 40, 31–4.

Remsen, J. V. and Ridgley, R. S. (1980). Additions to the avifauna of Bolivia. Condor, 82, 69–75.

Riddiford, N. (1986). Why do Cuckoos Cuculus canorus use so many species of hosts? Bird Study, 33, 1–5.

Remsen, J. V. and Traylor, M. A. (1983). Additions to the avifauna of Bolivia. Part 2. Condor, 85, 95–8.

Ridgley, R. S. and Greenfield, P. J. (2001). The birds of Ecuador. vols. 1, 2. Cornell University Press, Ithaca.

Remsen, J. V. and Traylor, M. A. (1989). An annotated list of the birds of Bolivia. Buteo, Vermilion, South Dakota.

Ridgway, R. (1912). Diagnoses of some new genera of American birds. Proceedings of the Biological Society of Washington, 25, 97–102.

Remsen, J. V., Traylor, M. A., and Parkes, K. C. (1986). Range extensions for some Bolivian birds, 2 (Columbidae to Rhinocryptidae). Bulletin of the British Ornithologists’ Club, 106, 22–32.

Ridgway, R. (1916). The birds of North and Middle America. Part VII. United States National Museum Bulletin 50.

Rensch, B. (1924). Zur Entstehung der Mimikry er Kuckuckseier. Journal für Ornithologie , 72, 471–2.

Riley, J. (1997). The birds of Sangihe and Talaud, north Sulawesi. Kukila, 9, 3–36.

Rensch, B. (1931). Die Vogelwelt von Lombok, Sumbawa und Flores (aus den Ergebnissen der SundaExpedition Rensch). Mitteilungen aus der Zoologischen Museum in Berlin, 17, 451–637.

Riley, J. H. (1924). A collection of birds from north and north-central Celebes. Proceedings of the United States National Museum, 64 (16), 1–18.

Rensch, B. (1950). Die Abhängigkeit der relativen Sexualdifferenz von der Körpergrösse. Bonner Zoologische Beitrage, 1, 58–69.

Riley, J. H. (1938). Birds from Siam and the Malay peninsula in the United States National Museum collected by Drs. Hugh M. Smith and William L. Abbott. United States National Museum Bulletin, 172.

Rea, A. M. (1983). Once a river. University of Arizona Press, Tucson.

Rinke, D. (1999). Zucht des Renauldkuckucks (Carpococcyx renauldi) im Vogelpark Walsrode. Voliere, 22, 212–4.

Rey, E. (1892). Altes und Neues aus dem Haushalte des Kuckucks. R. Freese, Leipzig.

Ripley, S. D. (1941). Notes on a collection of birds from northern Celebes. Occasional Papers of the Boston Society of Natural History, 8, 343–58.

Reynard, G. B. and Sutton, R. L. (2000). Bird songs in Jamaica. Library of Natural Sounds, Ithaca, New York. Rich, P.V. and Baird, R. F. (1986). History of the Australian avifauna. Current Ornithology, 4, 97–139. Richardson, C. (1990). The birds of the United Arab Emirates. Hobby, Dubai. Richmond, C. W. (1903). Birds collected byDr. W. L. Abbottt and Mr. C. B. Kloss in the Andaman and Nicobar Islands. Proceedings of the United States National Museum, 25, 287–314.

Ripley, S. D. (1942). Notes on Malaysian cuckoos. Auk, 59, 575–6. Ripley, S. D. (1943). Rhinortha chlorophaea in Borneo. Auk, 60, 604–5. Ripley, S. D. (1944). The bird fauna of the west Sumatra islands. Bulletin of the Museum of Comparative Zoölogy , 94, 307–430. Ripley, S. D. (1946). Comments on Ceylon birds. Spolia Zeylanica, 24, 197–241.

Bibliography 585 Ripley, S. D. (1950). Birds from Nepal 1947–1949. Journal of the Bombay Natural History Society, 49, 354–417. Ripley, S. D. (1951). Migrants and introduced species in the Palau Archipelago. Condor, 53, 299–300. Ripley, S. D. (1959). Birds from Djailolo, Halmahera. Postilla, 41, 1–8. Ripley, S. D. (1964). A systematic and ecological study of birds of New Guinea. Peabody Museum of Natural History Yale University Bulletin, 19. Ripley, S. D. and Beehler, B. M. (1989). Ornithogeographic affinities of the Andaman and Nicobar Islands. Journal of Biogeography, 16, 323–32. Ripley, S. D. and Bond, G. M. (1966). The birds of Socotra and Abd-el-Kuri. Smithsonian Miscellaneous Collections, 4 (7), 1–37. Ripley, S. D. and Heinrich, G. (1966a). Additions to the avifauna of northern Angola. Postilla, 95, 1–29.

Robertson, W. B. (1978). Mangrove Cuckoo. In Rare and endangered biota of Florida, vol. 4, (ed. H. W. Kale), pp. 57–8. University Press of Florida, Gainesville. Robiller, F., Michi, H., and Michi, M. (1992). Über den Renauldkuckuck (Carpococcyx renauldi Oustalet, 1896) und seine Zucht in der Forschungsstation Ornis Mallorca. Tropische Vögel , 13, 92–7. Robinson, H. C. and Kloss, C. B. (1919). On birds from South Annam and Cochin China. Part I. Ibis, Ser. 11, vol. 1, 392–453. Robinson, H. C. and Kloss, C. B. (1921–1924). The birds of south-west and peninsular Siam. Journal of the Natural History Society of Siam, 5, 1–397. Robinson, H. C. and Kloss, C. B. (1922). Birds from the One Fathom Bank Lighthouse, Straits of Malacca, November, 1918 and November and December, 1919. Journal of the Federated Malay States Museum, 10, 253–5.

Ripley, S. D. and Heinrich. G. (1966b). Comments on the avifauna of Tanzania I. Postilla, 96, 1– 45.

Robinson, H. C. and Kloss, C. B. (1923a). A nominal list of the birds of Sumatra, appendix and notes. Journal of the Federated Malay States Museum, 8, 319–62.

Ripley, S. D. and Rabor, D. S. (1958). Notes on a collection of birds from Mindoro Island, Philippines. Peabody Museum of Natural History,Yale, University Bulletin, 13, 1–83.

Robinson, H. C. and Kloss, C. B. (1923b). On a large collection of birds chiefly from West Sumatra made by Mr. E. Jacobson. Journal of the Federated Malay States Museum, 11, 189–347.

Robbins, C. S., Bystrak, D., and Geissler, P. H. (1986). The Breeding Bird Survey: its first fifteen years, 1965–1979. United States Fish and Wildlife Service Resource Publication, 157.

Robinson, H. C. and Kloss, C. B. (1931). Some birds from Siam and Laos (middle Mekong). Ibis, Ser. 13, vol. 1, 319–41.

Robert, M. and Sorci, G. (1999). Rapid increase of host defence against brood parasites in a recently parasitized area: the case of Village Weavers in Hispaniola. Proceedings of the Royal Society of London, Series B, 266, 941– 6. Robert, M. and Sorci, G. (2001). The evolution of obligate interspecific brood parasitism in birds. Behavioral Ecology, 12, 128–35.

Robinson, S. K. (1997). Birds of a Peruvian oxbow lake: populations, resources, predation, and social behavior. Ornithological Monographs, 48, 613–39. Robinson, S. K. and Terborgh, J. (1997). Bird community dynamics along primary successional gradients of an Amazonian whitewater river. Ornithological Monographs, 48, 641–72.

Roberts, A. (1940). The birds of South Africa. H. F. and G. Witherby, London.

Robinson, S. K., Fitzpatrick, J. W., and Terborgh, J. (1995). Distribution and habitat use of Neotropical migrant landbirds in the Amazon basin and Andes. Bird Conservation International, 5, 305–23.

Roberts, T. S. (1936). The birds of Minnesota, vol. 1. University of Minnesota Press, Minneapolis.

Robson, C. (2000a). A guide to the birds of southeast Asia. Princeton University Press, Princeton.

Roberts, T. J. (1991). The birds of Pakistan, vol. 1. Oxford University Press, Karachi.

Robson, C. (2000b). From the field. Oriental Bird Club Bulletin, 31, 49–57.

Robertson, A. and Jackson, M. C. A. (1992). Birds of Periyar. Tourism and Wildlife Society of India.

Robson, C. (2002). Birds of Thailand. Princeton University Press, Princeton.

586 Bibliography Robson, C. R., Eames, J. C., Nguyen Cu, and Trong Van La. (1993). Further recent records of birds from Viet Nam. Forktail, 8, 25–52. Robson, C. R., Buck, H., Farrow, D. S., Fischer, T., and King, B. F. (1998). A birdwatching visit to the Chin Hills, west Burma (Myanmar), with notes from nearby areas. Forktail, 13, 109–20. Roché, J. C. (1971). Birds of Madagascar (1 sound disk). J. C. Roché, Aubenas-les-Alpes. Roche, J. P. and Glanz, W. E. (1998). Nestling aggression and the evolution of brood parasitism in altricial birds. Ethology, Ecology and Evolution, 10, 287–92.

Røskaft, E., Moksnes, A., Meilvang, D., Bicik, V., Jemelikova, J. and Honza, M. (2002). No evidence for recognition errors in Acrocephalus warblers. Journal of Avian Biology, 33, 31–8. Ross, C. A. and Ramos, T. R. (1992). Distribution of Centropus viridis in the Babuyan Islands, northern Philippines. Bulletin of the British Ornithologists’ Club, 112, 180–1. Rossouw, J. and Sacchi, M. (1998). Where to watch birds in Uganda. Uganda Tourist Board, Kampala. Roth, P. (1981). A nest of the Rufous-vented Groundcuckoo (Neomorphus geoffroyi). Condor, 83, 388.

Rodner, C., Lentino, M., and Restall, R. (2000). Checklist of the birds of northern South America. Pica, Sussex.

Rothschild, M. and Clay, T. (1952). Fleas, flukes and cuckoos. New Naturalist, Collins, London.

Rodwell, S. P. (1996). Notes on the distribution and abundance of birds observed in Guinea-Bissau, 21 February to 3 April 1992. Malimbus, 18, 25–43.

Rothschild, W. (1921). On a collection of birds from west-central and north-western Yunnan. Novitates Zoologicae, 28, 14–26.

Rohling, E. J., Fenton, M., Jorissen, F. J., Bertrand, P., Ganssen, G., and Caulets, J. P. (1998). Magnitudes of sealevel lowstands of the past 500,000 years. Nature, 394, 162–5.

Rothschild, W. (1925). On a fourth collection of birds made by Mr. George Forrest in north-western Yunnan. Novitates Zoologicae, 32, 292–313.

Rohwer, S. and Spaw, C. D. (1988). Evolutionary lag versus bill-size constraints: a comparative study of the acceptance of cowbird eggs by old hosts. Evolutionary Ecology, 2, 27–36.

Rothschild, W. (1926). On the avifauna of Yunnan, with critical notes. Novitates Zoologicae, 33, 189–394. Rothschild, W. and Hartert, E. (1907). Notes on Papuan birds. Novitates Zoologicae, 14, 433– 46.

Rohwer, S., Spaw, C. D., and Røskaft, E. (1989). Costs to Northern Orioles of puncture-ejecting parasitic cowbird eggs from their nests. Auk, 106, 734–8.

Rothschild, W. and Hartert, E. (1914a). On a collection of birds from Goodenough Island. Novitates Zoologicae, 21, 1–9.

Roldan, E. R. S. and Gomendio, M. (1999). The Y chromosome as a battle ground for sexual selection. Trends in Ecology and Evolution, 14, 58– 62.

Rothschild, W. and Hartert, E. (1914b). On the birds of Rook Island, in the Bismarck Archipelago. Novitates Zoologicae, 21, 207–18.

Root, T. (1988). Atlas of wintering North American birds: an analysis of Christmas Bird Count data. University of Chicago Press, Chicago.

Rothschild, W., Stresemann, E., and Paludan, K. (1932). Ornithologische Ergebnisse der Expedition Stein 1931–32. I. Die Vögel von Waigeu; II. Die Vögel von Numfor; III. Die Vögel von Japen (= Jobi). Novitates Zoologicae, 38, 127–88, 188–207, 207–47.

Rose, L. N. (1982). Breeding ecology of British pipits and their Cuckoo parasite. Bird Study, 29, 27–40. Rosenberg, D. K., Wilson, M. H., and Cruz, F. (1990). The distribution and abundance of the Smooth-billed Ani Crotophaga ani (L.) in the Galapagos Islands, Ecuador. Biological Conservation, 51, 113–23. Røskaft, E., Rohwer, S., and Spaw, C. D. (1993). Cost of puncture ejection compared with costs of rearing cowbird chicks for Northern Orioles. Ornis Scandinavica, 24, 28–32.

Rothstein, S. I. (1975a). An experimental and teleonomic investigation of avian brood parasitism. Condor, 77, 250–71. Rothstein, S. I. (1975b). Evolutionary rates and host defenses against avian brood parasitism. American Naturalist, 109, 161–76. Rothstein, S. I. (1990). A model system for coevolution: avian brood parasitism. Annual Review of Ecology and Systematics, 21, 481–508.

Bibliography 587 Rothstein, S. I. and Robinson, S. K., eds. (1998). Parasitic birds and their hosts. Oxford University Press, Oxford. Round, P. D. (1988). Resident forest birds in Thailand: their status and conservation. ICBP Monograph no. 2.

Russell, S. M. (1980). Distribution and abundance of North American migrants in lowlands of northern Colombia. In Migrant birds in the Neotropics, (eds. A. Keast and E. S. Morton), pp. 249–52. Smithsonian Institution Press, Washington, D.C.

Round, P. D. (1995). Recent reports December 1994January 1995. Bird Conservation Society of Thailand Bulletin, 12, 14.

Russell, S. M. and Monson, G. (1998). The birds of Sonora. University of Arizona Press, Tucson.

Round, P. D. and U. Treesucon. (1997). Birds of Khao Nor Chuchi. Bird Conservation Society of Thailand, Bangkok.

Rütschke, E. (1966). Die submikroskopische Struktur schillernder Federn von Entenvögeln. Zeitschrift für Zellforschung, 73, 432–43.

Roux, J.-P. and Martinez, J. (1987). Rare, vagrant and introduced birds at Amsterdam and Saint Paul Islands, southern Indian Ocean. Cormorant, 14, 3–19. Rowan, M. K. (1983). The doves, parrots, louries and cuckoos of southern Africa. John Voelcker Bird Book Fund, Cape Town. Rowan, M. K. and Broekhuysen. G. J. (1962). A study of the Karoo Prinia. Ostrich, 33 (2), 6–30. Rowlands, B. W., Trueman, T., Olson, S. L., McCulloch, M. N., and Brooke, R. K. (1998). The birds of St Helena. BOU Checklist No. 16. British Ornithologists’ Union, Tring. Rowley, I. and Russell, E. (1997). Fairy-wrens and grasswrens. Oxford University Press, Oxford. Rowley, I., Brooker, M., and Russell, E. (1991). The breeding biology of the Splendid Fairy-wren Malurus splendens: the significance of multiple broods. Emu, 91, 197–221.

Saccone C., De Giorgi, C., Gissi, C., Pesole, G., and Reyes, A. (1999). Evolutionary genomics in metazoa: the mitochondrial DNA as a model system. Gene, 238, 195–209. Saha, B. C. and Dasgupta, J. M. (1992). Birds of Goa. Records of the Zoological Survey of India Occasional Paper no. 143, 1–56. Salaman, P., Cuadros, T., Jaramillo, J. G., and Weber, W. H. (2001). Lista de Chequeo de las aves de Colombia. Sociedad Antioqueña Ornitologia, Medellín, Colombia. Salewski, V. (2000). The birds of Comoé National Park, Ivory Coast. Malimbus, 22, 55–76. Salewski, V. and Grafe, T. U. (1999). New tape recordings of three West African birds. Malimbus, 21, 117–21. Salomonsen, F. (1953). Miscellaneous notes on Philippine birds. Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening i København , 115, 205–81.

Rowley, J. S. (1984). Breeding records of land birds in Oaxaca, Mexico. Proceedings of the Western Foundation of Vertebrate Zoology, 2 (3).

Salter, B. E. (1978). A note on the Channel-bill Cuckoo. Bird Observer, 559, 24.

Royston, A. (1981). Notes on two African species. Avicultural Magazine, 87, 111–2.

Salvador, S. A. (1981). Desarrollo de una nidada communal de Pirincio, Guira guira. Historia Natural (Mendoza, Corrientes, Argentina), 2, 29–31.

Royte, E. (2001). The tapir’s morning bath. Houghton Mifflin, Boston. Rozendaal, F. G. and Dekker, R. W. R. J. (1989). Annotated checklist of the birds of the Dumoga-Bone National Park, North Sulawesi. Kukila, 4, 85–109. Russell, E. M. (1989). Co-operative breeding—a Gondwanan perspective. Emu, 89, 61–2. Russell, E. M. and Rowley, I. (1993). Demography of the co-operatively breeding Splendid Fairy-wren Malurus splendens. Australian Journal of Zoology, 41, 475–505.

Salvador, S. A. (1982). Estudio de parasitismo del crespin Tapera naevia chochi (Vieillot) (Aves: Cuculidae). Historia Natural (Mendoza, Corrientes, Argentina), 2, 65–70. Salvadori, T. (1874). Catalogo sistematico degli Uccelli di Borneo con note ed osservazioni di G. Doria ed O. Beccaari intorno alle specie da essi raccolte nel Ragialto di Sarawak. Annali del Museo Civico di Storia Naturale di Genova, 5, 1–429. Salvadori, T. (1875). Descrizione di cinquantotto nuove specie di uccelli, ed osservazioni intorno ad altre poco note, della Nuova Guinea e di altre Isole Papuane,

588 Bibliography raccolte dal Dr. Odoardo Beccari e dai cacciatori del Sig. A. A. Bruijn. Annali del Museo Civico di Storia Naturale di Genova, 7, 896–992. Salvadori, T. (1876). Catalogo di una seconda collezione di uccelli raccolti del Sig. L. M. D’Albertis nell’ Isola Yule e sulla vicina costa della Nuova Guinea e di una piccola collezione della regione bagnata dl fiume Fly. Annali del Museo Civico di Storia Naturale di Genova, 9, part 2, 7–49. Salvadori, T. (1878a). Reports on the collection of birds made during the Voyage of H. M. S. ‘Challenger’. No. vi. On the birds of Ternate, Amboyna, Banda, the Ké islands, and the Aru Islands. Proceedings of the Zoological Society of London, 1878, 78–100. Salvadori, T. (1878b). Prodromus Ornithologiae Papuasiae et Moluccarum. VI. Picariae. Fam. Cuculidae. Annali del Museo Civico di Storia Naturale di Genova, 13, 456–63. Salvadori, T. (1879). Catalogo di una Collezione di Uccelli fatta nella parte occidentale di Sumatra dal Prof. Odoardo Beccari. Annali del Museo Civico di Storia Naturale di Genova, 14, 169–253. Salvadori, T. (1881). Ornitologia della Papuasia e delle Molucche. Parte prima. Memoire della Reale Accademia delle Scienze di Torino, (Serie II), 33, i-xii and 1–576. Salvadori, T. (1889). Aggiunte alla ornitologia della Papuasia e delle Molucche. Parte prima. Accipitres – Psittaci – Picariae. Memoires della Reale des Accademia della Scienze di Torino (2), 40, 131–92. Saunders, D. and Ingram, J. (1995). Birds of southwestern Australia. Surrey Beatty, Sydney. Saunders, G. B. (1961). Yellow-billed Cuckoo in stomach of tiger shark. Auk, 79, 118. Sayer, J. A., Harcourt, C. S., and Collins, N. M. (1992). The conservation atlas of tropical forests in Africa. Macmillan, London. Schäfer, E. and Meyer de Schauensee, R. (1939). Zoological results of the Second Dolan Expedition to Western China and Eastern Tibet, 1934–1936. Part 2, birds. Proceedings of the Academy of Natural Sciences of Philadelphia, 90, 185–260.

Schifter, H. (1973). A specimen of Coua delalandei (Temminck) (Cuculidae) in the Naturhistorische Museum, Vienna (Austria). Bulletin of the British Ornithologists’ Club, 93, 2–3. Schlegel, H. (1862). Museum d’Histoire Naturelle de PaysBas, Revue méthodique et critique de la collection des oiseaux. Leiden. Schmidl, D. (1982). The birds of the Serengeti National Park, Tanzania. British Ornithologists’ Union Check-list, no. 5. Schmitt, C. G., Schmitt, D. C. and Remsen, J. V. (1997). Birds of the Tambo area, an arid valley in the Bolivian Andes. Ornithological Monographs, 48, 701–16. Schmidt, W. J. (1964). Calcosphäriten in der Säulenlage der Guira-Eischale. Zoologischer Anzeiger, 172, 426–30. Schodde, R., Fullagar, P., and Hermes, N. (1983). A review of Norfolk Island birds: past and present. Australian National Parks and Wildlife Service Special Publication, 8. Schönwetter, M. (1964). Handbuch der Oologie. Band 1, Nonpasseres. Cuculiformes: Nos. 9, 10. Ed. W. Meise. Akademie-Verlag, Berlin. Schouteden, H. (1968). La faune Ornithologique du Kivu. 1. Non Passereaux. Zoologisch Documentaire 12. Schubert, M. (1982). Stimmen der Vögel Zentralasiens . 33 rpm recording, 8.22.575/576, Eterna, VEB Deutsche Schallplatten, Berlin. Schulze-Hagen, K. (1992). Parasitierung und Brutverluste durch Kuckuck (Cuculus canorus) bei Teichund Sumpfrohrsänger (Acrocephalus scirpaceus, A. palustris) in Mittel- und West-Europa. Journal für Ornithologie , 133, 237–49. Schuster, L. (1926). Beiträge zu Verbreitung und Biologie der Vögel Deutsch-Ostafrikas, Teil 2. Journal für Ornithologie, 74, 521–41. Schwartz, A. and Klihikowski, R. F. (1965). Additional observations on West Indian birds. Notulae Naturae, 376, 1–16.

Scharnke, H. (1930). Physiologisch-anatomische Studien am Fuß der Spechte. Journal für Ornithologie , 78, 308–27.

Sclater, P. L. and Hudson, W. H. (1888). Argentine ornithology, a descriptive catalogue of the birds of the Argentine Republic, vol. 1. R. H. Porter, London.

Scharringa, J. (1999). Birds of tropical Asia, sounds and sights CD-ROM. Bird Songs International, Westernieland, The Netherlands.

Sclater, P. L. and Hudson, W. H. (1889). Argentine ornithology, a descriptive catalogue of the birds of the Argentine Republic, vol. 2. R. H. Porter, London.

Bibliography 589 Sclater, W. L. (1930). Systema avium aethiopicarum. British Ornithologists’ Union, London. Sclater, W. L. and Mackworth-Praed, C. (1919). A list of the birds of the Anglo-Egyptian Sudan, based on the collections of Mr. A. L. Butler, Mr. A. Chapman and Capt. H. Lynes, R. N., and Major Cuthbert Christy, R. A. M. C. (T. F.). Part III. Ibis, Ser. 11, vol. 9, 628–707. Sclater, W. L. and Moreau, R. E. (1932). Taxonomic and field notes on some birds of north-eastern Tanganyika Territory, part 1. Ibis, Ser. 13, vol. 2, 487–522. Scott, K. M. (1985). Allometric trends and locomotor adaptations in the Bovidae. Bulletin of the American Museum of Natural History, 179, 197–288. Sealy, S. G. (1978). Possible influence of food on egglaying and clutch size in the Black-billed Cuckoo. Condor, 80, 103–4. Sealy, S. G. (1985). Erect posture of the young Blackbilled Cuckoo: an adaptation for mobility in a nomadic species. Auk, 102, 889–92. Sealy, S. G. (2003). Laying times and a case of conspecific nest parasitism in the Black-billed Cuckoo. Journal of Field Ornithology, 74, 257–60. Sealy, S. G., McMaser, D. G., and Peer, B. D. (2002). Tactics of obligate brood parasites to secure suitable incubators. In Avian incubation: Behaviour, environment and evolution, (ed. D. C. Deeming), pp. 259–69. Oxford University Press, Oxford. Searcy, W. A. (1992). Measuring responses of female birds to male song. In Playback and studies of animal communication, (ed. P. K. McGregor), pp. 175–89. Plenum Press, New York. Seaton, C. (1962). The Yellow-breasted Sunbird as a host to the Rufous-breasted Bronze-cuckoo. Emu, 62, 174–6. Seebohm, H. (1882). Siberia in Asia: a visit to the valley of the Yenesay in East Siberia. John Murray, London. Seel, D. C. (1973). Egg-laying by the Cuckoo. British Birds, 66, 528–35.

Seel, D. C. (1984a). Geographical distribution of the Cuckoo Cuculus canorus in the western Palaearctic and Afro-tropical regions. Koninklijk Museum voor MiddenAfrika, Zoologische Wetenschappen, Annalen Series 8, Sciences Zoologiques, No. 239, 1–44. Seel, D. C. (1984b). Moult and body weight in the Cuckoo Cuculus canorus in the western Palaearctic and Afro-tropical regions. Koninklijk Museum voor MiddenAfrika, Zoologische Wetenschappen, Annalen Series 8, Sciences Zoologiques, No. 239, 49–91. Seel, D. C. and Davis, P. R. K. (1981). Cuckoos reared by unusual hosts in Britain. Bird Study, 28, 242–3. Seel, D. C., Walton, K. G., and Wyllie, I. (1981). Age of first breeding in the Cuckoo. Bird Study, 28, 211–4. Sefc, K., Payne, R. B., and Sorenson, M. D. (2003). Phylogenetic relationships of African sunbird-like warblers: Moho Hypergerus atriceps, Green Hylia Hylia prasina and Tit-Hylia Pholidornis rushiae. Ostrich, 74, 8–17. Seger, J. and Brockmann, H. J. (1987). What is bethedging? Oxford Surveys in Evolutionary Biology, 4, 182–211. Seibel, D. E. (1988). A phylogenetic analysis of the Cuculiformes and Opisthocomus, based on postcranial skeletal characters. PhD dissertation, University of Kansas, Lawrence, Kansas. UMI 8903159. Selander, R. K. (1965). On mating systems and sexual selection. American Naturalist, 99, 129–41. Selander, R. and Giller, D. R. (1959). The avifauna of the Barranca de Orlatos, Jalisco, Mexico. Condor, 61, 210–22. Selous, F. C. (1912–13). Early laying of the Cuckoo and removal of the eggs of foster-parents. British Birds, 6, 90–1. Seppä, J. (1969). The Cuckoo’s ability to find a nest where it can lay an egg. Ornis Fennica, 46, 78–9. Serie, P. (1923a). Un huevo de pirincho Guira guira en un nido de Phytotoma rutila. Hornero, 3, 100. Serie, P. (1923b). Huevos de pirincho en nido de chimango. Hornero, 3, 189.

Seel, D. C. (1977a). Migration of the northwestern European population of the Cuckoo Cuculus canorus as shown by ringing. Ibis, 119, 309–22.

Serle, W. (1950a). Notes on the birds of south-western Nigeria. Ibis, 92, 84–94.

Seel, D. C. (1977b). Trapping season and body size in the Cuckoo. Bird Study, 24, 114–8.

Serle, W. (1950b). A contribution to the ornithology of the British Cameroons. Ibis, 92, 343–76.

590 Bibliography Serle, W. (1954). A second contribution to the ornithology of the British Cameroons. Ibis, 96, 47–80.

Sharpe, R. B. (1879). A list of the birds of Labuan Island and its dependencies. Proceedings of the Zoological Society of London, 1879, 317–55.

Serle, W. (1965). A third contribution to the ornithology of the British Cameroons. Ibis, 107, 60–94.

Sharpe, R. B. (1884). Notes on a collection of birds made by Herr F. Bohndorff in the Bahr el Ghazal Province and the Nyam-Nyam Country in equatorial Africa. Journal of the Linnean Society, Zoology, 17, 419–41.

Serle, W. (1977). The aberrant eggs of Turdoides plebejus in Nigeria and their relation to cooperative breeding and to victimisation by Clamator cuckoos. Bulletin of the British Ornithologists’ Club, 97, 39–41. Serle, W. and Morel, G. (1977). A field guide to the birds of West Africa. Collins, London. Serventy, D. L. and Whittell, H. M. (1976). Birds of Western Australia. (5th edn.) University of Western Australia Press, Perth. Severinghaus, S. R. and Blackshaw, K. T. (1976). The birds of Taiwan. Mei Ya Publications, Taipei. Sganzin, V. (1840). Notes sur les mammifères et sur l’ornithologie de l’île de Madagascar (1831 et 1832). Mémoires de la Société du Museum d’histoire naturelle de Strasbourg, 3 (1), article 3, 1–49. Shanahan, M., So, S., Compton, S. G., and Corlett, R. (2001). Fig-eating by vertebrate frugivores: a global review. Biological Reviews, 76, 531–72. Shapiro, B., Sibthorpe, D., Rambaut, A., Austin, J., G. M. Wragg, G. M., Bininda-Edmonds, O. R. P., Lee, P. L. M., and Cooper, A. 2002. Flight of the Dodo. Science, 295, 1683. Sharland, R. E. and Serle, W. (1977). The egg colour of Clamator levaillantii (Swainson) in Nigeria. Bulletin of the Nigerian Ornithologists’ Society, 13, 80. Sharp, A. (1976). Indian Mynahs as hosts to Great Spotted Cuckoo. Natal Bird Club Newsletter, 242, 3–4. Sharpe, C. (1987). Barred Long-tailed Cuckoo and African Broadbill in the South-east Lowveld. Honeyguide, 33, 150. Sharpe, R. B. (1873). On the Cuculidae of the Ethiopian Region. Proceedings of the Zoological Society of London, 1873, 578–624.

Sharpe, R. B. (1890). On the ornithology of Northern Borneo, with notes by John Whitehead, part 5. Ibis, 6th Ser, vol. 2, 1–24. Shaw, T. H. (1938). Nordchinesische Blauelstern als Pflegeeltern von Cuculus micropterus. Ornithologische Monatsberichte, 46, 154–5. Sheehan, D. (1999). Winter calling by Klaas’s Cuckoo. Honeyguide, 45, 136–7. Sheldon, F. H., Moyle, R. G., and Kennard, J. (2001). Ornithology of Sabah: history, gazetteer, annotated checklist, and bibliography. Ornithological Monographs, 52, 1–285. Shelford, R. (1900). On the pterylosis of the embryos and nestlings of Centropus sinensis. Ibis, Ser. 6, vol. 24, 654–67. Shelford, R. W. C. (1916). A naturalist in Borneo, ed. E. B. Poulton. T. Fisher Unwin, London. Shelley, G. E. (1891). Cuculidae. In Catalogue of the birds in the British Museum, 19, (eds. W. L. Sclater and G. E. Shelley), pp. 209–434. British Museum (Natural History), London. Sherry, D. F., Forbes, M. R. L., Khurgel, M., and Ivy, G. O. (1993). Females have a larger hippocampus than males in the brood-parasitic Brown-headed Cowbird. Proceedings of the National Academy of Sciences USA, 90, 7839–43. Shetlar, D. J. (1971). Winter food of a central Oklahoma Roadrunner. Bulletin of the Oklahoma Ornithological Society, 4, 35. Shirihai, H. (1996). The birds of Israel. Academic Press, London.

Sharpe, R. B. (1876). Contributions to the ornithology of Borneo. Part 1. Ibis, 3rd Ser., vol. 4, 29–52.

Short, L. L. (1974). Nesting of southern Sonoran birds during the summer rainy season. Condor, 76, 21–32.

Sharpe, R. B. (1878). Contributions to the ornithology of New Guinea, part 1. Journal of the Linnean Society, Zoology, 13, 79–83.

Short, L. L. (1975). A zoogeographic analysis of the South American chaco avifauna. Bulletin of the American Museum of Natural History, 154, 163–352.

Bibliography 591 Short, L. L. (1976). Notes on a collection of birds from the Paraguayan chaco. American Museum Novitates, 2597, 1–16.

Sick, H. (1953a). Zur Kenntnis der brasilianischen Lerchenkuckucke Tapera und Dromococcyx. Bonner Zoologische Beiträge , 4, 305–26.

Short, L. L. and Horne, J. F. M. (2001). Toucans, barbets and honeyguides. Oxford University Press, Oxford.

Sick, H. (1953b). Anotações sôbre cucos Brasileiros (Cuculidae, Aves). Revista Brasileira de Biologia, 13 (2), 145–68.

Short, L. L., Horne, J. F. M., and Muringo-Gichuki, C. (1990). Annotated check-list of the birds of East Africa. Proceedings of the Western Foundation of Vertebrate Zoology, 4, 61–246.

Sick, H. (1981). Zum Problem der Elimination der Nestgeschwister beim Lerchenkuckuck Tapera naevia. Journal für Ornithologie , 122, 437–8.

Shufeldt, R. W. (1886a). The skeleton in Geococcyx. Journal of Anatomy and Physiology, 20, 244–65.

Sick, H. (1993). Birds in Brazil. Princeton University Press, Princeton.

Shufeldt, R. W. (1886b). Contributions to the anatomy of Geococcyx californianus. Proceedings of the Zoological Society of London, 1886 (1887), 466–91.

Sick, H. (1997). Ornitologia brasileira. Editora Nova Fronteira, Rio de Janeiro.

Shufeldt, R. W. (1886c). Osteological note upon the young of Geococcyx californianus. Journal of Anatomy and Physiology, 21, 101–2.

Siebers, H. C. (1930). Fauna Buruana; Aves. Treubia, 7, Supplement, 5, 165–303.

Shufeldt, R. W. (1901). The osteology of the cuckoos (Coccyges). Proceedings of the American Philosophical Society, 40, 4–51.

Siegel, C. E., Hamilton, J. M., and Rogerio Castro, N. (1989). Observations of the Red-billed Ground-cuckoo (Neomorphus pucheranii) in association with tamarins (Saguinas) in northeastern Amazonian Peru. Condor, 91, 720–2.

Shufeldt, R. W. (1915). Fossil birds in the Marsh Collection of Yale University. Transactions of the Connecticut Academy of Arts and Sciences, 19, 1–109.

Sieving, K. E. (1990). Pheasant Cuckoo foraging behavior, with notes on habits and possible social organization in Panama. Journal of Field Ornithology, 61, 41–6.

Sibley, C. G. (1951). Notes on the birds of New Georgia, central Solomon Islands. Condor, 53, 81–92.

da Silva, J. M. C., Oren, D. C., Roma, J. C., and Henriques, L. M. P. (1997). Composition and distribution patterns of the avifauna of an Amazonian upland savanna, Amapá, Brazil. Ornithological Monographs, 48, 743–62.

Sibley, C. G. (1955). A synopsis of the birds of the world. C. G. Sibley, Ithaca, New York. Sibley, C. G. and Ahlquist, J. E. (1972). A comparative study of the egg white proteins of non-passerine birds. Peabody Museum of Natural History,Yale University, Bulletin 39.

Simpson, S. and Cracraft, J. (1981). The phylogenetic relationships of the Piciformes (Class Aves). Auk, 98, 481–94.

Sibley, C. G. and Ahlquist, J. E. (1973). The relationships of the Hoatzin. Auk, 90, 1–13.

Simpson, K. and Day, N. (1996). The Princeton field guide to the birds of Australia. (5th edn.) Princeton University Press, Princeton.

Sibley, C. G. and Ahlquist, J. E. (1990). Phylogeny and classification of birds: a study in molecular evolution. Yale University Press, New Haven.

Singh, P. (1994). Recent bird records from Arunachal Pradesh. Forktail, 10, 65–104.

Sibley, C. G. and Monroe, B. L. (1990). Distribution and taxonomy of birds of the world. Yale University Press, New Haven. Sick, H. (1949). Beobachtungen an dem brasilianischen Bodenkuckuck Neomorphus geoffroyi dulcis Snethlage. In Ornithologie als Biologische Wissenschaft, (eds. E. Mayr and E. Schutz), pp. 229–39. Heidelberg, Carl Winter.

Sivakumar, K. and Sankaran, R. (2002). New records of birds from the Andaman and Nicobar Islands. Forktail, 18, 149–50. Skead, C. J. (1946). Record of a young Black Cuckoo (Surniculoides clamosus). Ostrich, 17, 359–60. Skead, C. J. (1947). A study of the Cape Weaver. Ostrich, 18, 1–42.

592 Bibliography Skead, C. J. (1951). Cuckoo studies on a South African farm (part I). Ostrich, 22, 163–75.

Slud, P. (1967). The birds of Cocos Island. Bulletin of the American Museum of Natural History, 134, 261–96.

Skead, C. J. (1952). Cuckoo studies on a South African farm (part II). Ostrich, 23, 2–15.

Smedley, D. I. (1983). Mating of the Fan-tailed Cuckoo. Corella, 7, 43.

Skead, C. J. (1954). A study of the Larger Double-collared Sunbird (Cinnyris afra afra Linnaeus). Ostrich, 25, 76–88.

Smetacek, V. (1975). On the increasing occurrence of typically plains-birds in the Kumaon Hills. Journal of the Bombay Natural History Society, 71, 299–302.

Skead, C. J. (1967). Ecology of the birds in the Eastern Cape Province. Ostrich, Supplement, 7. Skead, C. J. (1971). A juvenile Klaas’s Cuckoo, Chrysococcyx klaas with its hosts in late June 1971. Ostrich, 43, 134. Skead, C. J. (1995). Life-history notes on East Cape bird species (1940–1990) . Vols. 1, 2. Algoa Regional Services Council, Port Elizabeth. Skinner, N. J. (1978). Notes on breeding of the Redchested Cuckoo in Nairobi. Scopus, 2, 59–62. Skinner, N. J. (1996). The breeding seasons of birds in Botswana II. Non-passerine families (sandgrouse to woodpeckers). Babbler, 31, 6–16. Skutch, A. F. (1950). The nesting season of central American birds in relation to climate and food supply. Ibis, 92, 185–222. Skutch, A. F. (1959). Life history of the Groove-billed Ani. Auk, 76, 281–317. Skutch, A. F. (1966). Life history notes on three tropical American cuckoos. Wilson Bulletin, 78, 139–65. Skutch, A. F. (1983). Birds of tropical America. University of Texas Press, Austin. Skutch, A. F. (1999). Trogons, Laughing Falcons, and other Neotropical birds. Texas A&M University Press, College Station. Slater, P. (1971). A field guide to Australian birds, nonpasserines. Livingston, Wynnewood, Pennsylvania. Slater, P., Slater, P., and Slater, R. (1986). The Slater field guide to Australian birds. Rigby, Dee Why West, New South Wales. Slatkin, M. (1974). Hedging one’s evolutionary bets. Nature, 250, 704–5. Slobodkin, L. B. (2001). The good, the bad and the reified. Evolutionary Ecology Research, 3, 1–13. Slud, P. (1964). The birds of Costa Rica, distribution and ecology. Bulletin of the American Museum of Natural History, 128.

Smith, H. C., Garthwaite, P. F., and Smythies, B. E. (1940). Notes on birds of Nattaung, Karenni. Journal of the Bombay Natural History Society, 41, 577–93. Smith, H. C., Garthwaite, P. F., Smythies, B. E., and Ticehurst, C. B. (1943). On the birds of the Karen hills and Karenni found over 3,000 feet. Journal of the Bombay Natural History Society, 44, 221–32. Smith, L. H. (1989). Feeding of young Pallid Cuckoo by four passerine species. Australian Bird Watcher, 13, 99–100. Smith, M. H. (1981). Breeding the Roadrunner Geococcyx californiana at Columbia Zoo. International Zoo Yearbook, 21, 119–21. Smith, S. and Hosking, E. (1955). Birds fighting. Experimental studies of the aggressive displays of some birds. Faber and Faber, London. Smith, S. M. (1971). The relationship of grazing cattle to foraging rates in anis. Auk, 88, 876–80. Smith, S. W. (1993a). Bird recordings from Sulawesi. cassette, the author, Lymington, U.K. Smith, S. W. (1993b). Bird recordings from the Moluccas. cassette, the author, Lymington, U.K. Smith, T. E. H. (1950). Black Drongos fostering a Koel. Journal of the Bombay Natural History Society, 49, 304. Smith, W. J. and Smith, A. M. (2000). Information about behavior is provided by songs of the Striped Cuckoo. Wilson Bulletin, 112, 491–97. Smithers, C. N. (1977). An instance of one adult Fan-tailed Cuckoo feeding another. Australian Birds, 12, 8. Smythies, B. E. (1940). The birds of Burma. American Baptist Mission Press, Rangoon. Smythies, B. E. (1949). A reconnaissance of the N’mai Hka drainage, northern Burma. Ibis, 91, 627–48. Smythies, B. E. (1953). The birds of Burma. (2nd edn.) Oliver and Boyd, Edinburgh.

Bibliography 593 Smythies, B. E. (1957). An annotated checklist of the birds of Borneo. Sarawak Museum Journal, vol. 9, no. 7, 523–818.

Soler, M. (1990). Relationships between the Great Spotted Cuckoo Clamator glandarius and its corvid hosts in a recently colonized area. Ornis Scandinavica, 21, 212–23.

Smythies, B. E. (1959). Bird notes from Mt Kinabalu. Sarawak Museum Journal, 9, 157–262.

Soler, M. and Martínez, J. G. (2000). Is egg-damaging behavior by Great Spotted Cuckoos an accident or an adaptation? Behavioral Ecology, 11, 495–501.

Smythies, B. E. (1960). Subspecific variation in birds’ songs and call-notes. Ibis, 102, 134–5. Smythies, B. (1986). The birds of Burma. (3rd edn.) Nimrod Press, Liss, U.K.

Soler, M. and Møller, A. P. (1990). Duration of sympatry and coevolution between the Great Spotted Cuckoo and its Magpie host. Nature, 343, 748–50.

Smythies, B. E. (1999). The birds of Borneo. (4th edn.) (revised by G. W. H. Davison). Natural History Publications (Borneo), Kota Kinabalu.

Soler, M. and Soler, J. J. (1991). Growth and development of Great Spotted Cuckoos and their Magpie hosts. Condor, 93, 49–54.

Snethlage, E. (1927). Ein neuer Cuculidae aus Südbrasilien. Ornitologische Monatsberichte, 36, 80–2.

Soler, M. and Soler, J. J. (1999). Innate versus learned recognition of conspecifics in Great Spotted Cuckoos. Animal Cognition, 2, 97–102.

Snethlage, H. (1928). Meine Reise durch Nordostbrasilien. III. Bausteine zur Biologie der angetroffenen Arten. Journal für Ornithologie , 76, 668–738. Snow, D. W. (1962). A field study of the Black and White Manakin, Manacus manacus, in Trinidad, W. I. Zoologica, 47, 99–221. Snow, D. W. (1978). An atlas of speciation in African nonpasserine birds. Trustees of the British Museum (Natural History), London. Snow, D. W. (1992). The Cuckoo controversy. In Birds, discovery and conservation, 100 years of the Bulletin of the British Ornithologists’ Club, (ed. D. W. Snow), pp. 81–6. Helm Information Ltd., Mountfield, U.K. Snow, D. W. and Snow, B. K. (1964). Breeding seasons and annual cycles of Trinidad land-birds. Zoologica (New York), 49, 1–39. Snyder, D. E. (1966). The birds of Guyana. Peabody Museum, Salem. Sody, H. J. V. (1989). Diets of Javanese birds. In Henri Jacob Victor Sody (1892–1959), his life and work , (ed. J. H. Becking), pp. 164–221. E. J. Brill, Leiden.

Soler, M., Palamino, J. J., Martínez, J. G., and Soler, J. J. (1994a). Activity, survival, independence and migration of fledgling Great Spotted Cuckoos. Condor, 96, 802–5. Soler, M., Soler, J. J., Martínez, J. G., and Møller, A. P. (1994b). Micro-evolutionary change in host response to a brood parasite. Behavioral Ecology and Sociobiology, 35, 295–301. Soler, M., Soler, J. J., Martínez, J. G., and Møller, A. P. (1995). Chick recognition and acceptance: a weakness in Magpies exploited by the parasitic Great Spotted Cuckoo. Behavioral Ecology and Sociobiology, 37, 243–8. Soler, M., Martínez, J. G., and Soler, J. J. (1996). Effects of brood parasitism by the Great Spotted Cuckoo on the breeding success of the Magpie host: an experimental study. Ardeola, 43, 87–96. Soler, M., Soler, J. J., and Martínez, J. G. (1997). Great Spotted Cuckoos improve their reproductive success by damaging Magpie host eggs. Animal Behaviour, 54, 1227–33.

Soler, J. J. and Møller, A. P. (1996). A comparative analysis of the evolution of variation in appearance of eggs of European passerines in relation to brood parasitism. Behavioral Ecology, 7, 89–94.

Soler, M., Soler, J. J., and Martínez, J. G. (1998). Duration of sympatry and coevolution between the Great Spotted Cuckoo Clamator glandarius and its primary host, the Magpie Pica pica. In Parasitic birds and their hosts, (eds. S. I. Rothstein and S. K. Robinson), pp. 113–28. Oxford University Press, Oxford.

Soler, J. J., Møller, A. P., and Soler, M. (1999). A comparative study of host selection in the European Cuckoo (Cuculus canorus). Oecologia, 118, 265–76.

Soler, M., Soler, J. J., and Møller, A. P. (2000). Effect of the Great Spotted Cuckoo presence on Magpie rejection behaviour. Behaviour, 137, 213–20.

594 Bibliography Sorenson, M. D. (1991). The functional significance of parasitic egg laying and typical nesting in Redhead ducks: an analysis of individual behaviour. Animal Behaviour, 42, 771–96. Sorenson, M. D. (1993). Parasitic egg laying in Canvasbacks: frequency, success, and individual behavior. Auk, 110, 57–69. Sorenson, M. D. (1997). Effects of intra- and interspecific brood parasitism on a precocial host, the Canvasback, Aythya valisineria. Behavioral Ecology, 8, 153–61. Sorenson, M. D. and Payne, R. B. (2001). A single, ancient origin of obligate brood parasitism in African finches: implications for host-parasite coevolution. Evolution, 55, 2550–67. Sorenson, M. D. and Payne, R. B. (2002). Molecular genetic perspectives on avian brood parasitism. Integrative and Comparative Biology, 42, 388–400. Sorenson, M. D., Ast, J. C., Dimcheff, D. E., Yuri, T., and Mindell, D. P. (1999). Primers for a PCR-based approach to mitochondrial genome sequencing in birds and other vertebrates. Molecular Phylogenetics and Evolution, 12, 105–14. Sorenson, M. D., Sefc, K. M., and Payne, R. B. (2003a). Speciation by host switch in brood parasitic indigobirds. Nature, 424, 928–31. Sorenson, M. D., Oneal. E., García-Moreno, J., and Mindell, D. P. (2003b). More taxa, more characters: the Hoatzin problem is still unresolved. Molecular Biology and Evolution, 20, 1484–99. Southern, H. N. (1954). Mimicry in cuckoos’ eggs. In Evolution as a process, (eds. J. Huxley, A. C. Hardy and E. B. Ford), pp. 257–70. Collier, New York. Souza, F. L. (1995). A study of group structure and home range size of Crotophaga ani and Guira guira in São Paulo, Brasil (Cuculiformes: Cuculidae). Ararajuba, 3, 72–4. Spencer, O. R. (1943). Nesting habits of the Blackbilled Cuckoo. Wilson Bulletin, 55, 11–22. Spennemann, A. (1928). Zur Brutbiologie von Centropus javanicus (Dumont). Beiträge zur Fortpflanzungsbiologie der Vögel , 4, 139–44. Spriggs, M. (1997). The island Melanesians. Blackwell, Cambridge, Massachusetts. Squelch, P. and Safe-Squelch, W. (1994). Didric Cuckoo being fed by Little Rush Warblers. East African Natural History Society Bulletin, 24, 39–40.

Squires, S. K. (1930). The Black-billed Cuckoo and tent caterpillar. Canadian Field Naturalist, 44, 185–7. Stager, K. E. (1964). The birds of Clipperton Island, eastern Pacific. Condor, 66, 357–71. Stahl, J. C., Weimerskirch, H., and Ridoux, V. (1984). Observations recentes d’oiseaux marins et terrestres visiteurs dans les îles Crozet, sud-ouest de l’Ocean Indien. Gerfaut, 74, 39–46. Stanford, J. K. (1931). The birds of the Prome District of Lower Burma, part III. Journal of the Bombay Natural History Society, 35, 32–50. Stanford, J. K. and Mayr, E. (1941). The Vernay-Cutting Expedition to northern Burma, part 5. Ibis, Ser. 14, vol. 5, 479–518. Stanford, J. K. and Ticehurst, C. B. (1935). Notes on some new or rarely recorded Burmese birds, part 2. Ibis, Ser. 13, vol. 5, 1–45. Stanford, J. K. and Ticehurst, C. B. (1939). On the birds of northern Burma, part V. Ibis, Ser. 14, vol. 3, 1–45. Stannard, J. (1966). Bird songs of Amanzi. Percy Fitzpatrick Institute of African Ornithology, Cape Town. Stark, A. C. and Sclater, W. L. (1903). The birds of South Africa, vol. 3. R. H. Porter, London. Starck, J. M. and Ricklefs, R. E., Eds. (1998). Avian growth and development. Oxford University Press, Oxford. Staub, F. (1976). Birds of the Mascarenes and Saint Brandon. Organisation Normale des Entreprises Ltée, Port Louis, Mauritius. Steadman, D. W. (1995). Prehistoric extinctions of Pacific island birds: biodiversity meets zooarchaeology. Science, 267, 1123–31. Steadman, D. W. and Kirch, P. V. (1998). Biogeography and prehistoric exploitation of birds in the Mussa Islands, Bismarck Archipelago, Papua New Guinea. Emu, 98, 13–22. Steadman, D. W. and Olson, S. L. (1985). Bird remains from an archaeological site on Henderson Island, South Pacific: man-caused extinctions on an “uninhabited” island. Proceedings of the National Academy of Sciences USA, 82, 6191–5. Steadman, D. W., White, J. P., and Allen, J. (1999). Prehistoric birds from New Ireland, Papua New Guinea: extinctions on a large Melanesian island. Proceedings of the National Academy of Sciences USA, 96, 2563–8.

Bibliography 595 Steadman, D. W., Plourde, A., and Burley, D. V. (2002). Prehistoric butchery and consumption of birds in the Kingdom of Tonga, South Pacific. Journal of Archaeological Science, 20, 571–84. Stearns, S. C. (1992). The evolution of life histories. Oxford University Press, Oxford. Stein, G. H. W. (1936). Ornithologische Ergebnisse der Expedition Stein 1931–1932. Journal fur Ornithologie, 84, 21–57. Steinbacher, J. (1935). Funktionell-anatomische Untersuchungen an Vogelfüßen mit Wendezehen und Rückzehen. Journal für Ornithologie , 83, 214–82. Stephan, B. (1970). Eutaxie, Diastataxie und andere Probleme der Befiederung des Vogelflügels. Mitteilungen aus dem Zoologischen Museum in Berlin, 46, 339–437. Stephan, B. (1991). Ontogenetischer und phylogenetischer Wandel des Modus der Handschwingenmauser. Mitteilungen aus dem Zoologischen Museum in Berlin, 67, Annalen für Ornithologie , 15, 121–35. Stephan, B. (2001a). Studien zur Morphologie der Kuckucke (Aves: Cuculiformes: Cuculidae). Teil I. Flügel: Form und Formwandel, Flügelformel, Maße, Anzahl der Hand- und Armschwinge, Alula. Zoologische Abhandlungen Staatliches Museum für Dresden , 51, 139–88. Stephan, B. (2001b). Studien zur Morphologie der Kuckucke (Aves: Cuculiformes: Cuculidae). Teil II. Schwanz; Maße, Proportionen, Stufung. Zoologische Abhandlungen Staatliches Museum für Dresden , 51, 377–401. Stettenheim, P. (1973). The bristles of birds. Living Bird, 12, 201–34. Stevens, H. (1915). Notes on the birds of Upper Assam, part 2. Journal of the Bombay Natural History Society, 23, 547–70. Stevens, H. (1924). Notes on the birds of the Sikkim Himalayas, part 3. Journal of the Bombay Natural History Society, 29, 1007–30. Stevens, H. (1925). Notes on the birds of the Sikkim Himalayas, part 6. Journal of the Bombay Natural History Society, 30, 664–85.

Stewart, R. E. and Robbins, C. S. (1958). Birds of Maryland and the District of Columbia. North American Fauna no. 62. Steyn, P. (1970). Journey for birds. Bokmakierie, 22, 86–92. Steyn, P. (1972). The development of Senegal Coucal nestlings. Ostrich, 43, 56–9. Steyn, P. (1973). Some notes on the breeding biology of the Striped Cuckoo. Ostrich, 44, 163–9. Steyn, P. (1996). Nesting birds, the breeding habits of southern African birds. Fernwood Press, Vlaeberg. Steyn, P. and Howells, W. W. (1975). Supplementary notes on the breeding biology of the Striped Cuckoo. Ostrich, 46, 258–60. Stiles, F. G. and Skutch, A. F. (1989). A guide to the birds of Costa Rica. Comstock, Ithaca, New York. Stiles, F. G., Telleria, J. L., and Diaz, M. (1995). Observaciones sobre la composicion, ecologia, y zoogeografia de la avifauna de la Sierra de Chiribiquete, Caqueta, Colombia. Caldasia, 17, 481–500. Stjernstedt, R. (1970). Birds in Brachystegia microphyllum in southern Tanzania. Bulletin of the British Ornithologists’ Club, 90, 28–31. Stjernstedt, R. (1984). First record of the Barred Longtailed Cuckoo Cercococcyx montanus in Zambia. Bulletion of the Zambian Ornithological Society, 16, 18–20. Stjernstedt, R. (1993). Birdsong of Zambia, 3 tape cassettes. the author, Powys, U.K. Stjernstedt, R. (1994). Rare birds of Zambia, tape cassette. the author, Powys, U.K. Stjernstedt, R. (1998). Courtship feeding and other observations on Klaas’s Cuckoo. Honeyguide, 44, 204–7. Stockton de Dod, A. (1978). Aves de la Republica Dominicana. Museo Nacional de Historia Natural, Santo Domingo, Republica Dominicana. Stockton de Dod, A. (1981). Guia de campo para las aves de la Republica Dominicana. Editora Horizontes de America, Santo Domingo, Republica Dominicana.

Stevenson, H. M. and Anderson, B. H. (1994). The birdlife of Florida. University Press of Florida, Gainesville.

Stoddart, D. R. (1984). Impact of man in the Seychelles. In Biography and Ecology of the Seychelles Islands, (ed. D. R. Stoddart). Monographiae Biologicae 55, 641–54, Dr. W. Junk, The Hague.

Stevenson, T. and Fanshawe, J. (2002). Field guide to the birds of East Africa. T. and A.D. Poyser, London.

Stokke, B. G., Moksnes, A., Røskaft, E., Rudolfsen, G., and Honza, M. (1999). Rejection of artificial Cuckoo

596 Bibliography (Cuculus canorus) eggs in relation to variation in egg appearance among Reed Warblers (Acrocephalus scirpaceus). Proceedings of the Royal Society of London, Series B, 266, 1483–8. Stone, W. (1908). A review of the genus Piaya Lesson. Proceedings of the Academy of Natural Sciences of Philadelphia, Oct. 1908, 492–501. Stone, W. (1929). On a collection of birds from the Pará region, eastern Brazil. Proceedings of the Academy of Natural Sciences of Philadelphia, 80, 149–76. Stone, W. (1933). Zoological results of the Dolan West China Expedition of 1931, part 1. Birds. Proceedings of the Academy of Natural Sciences of Philadelphia, 85, 165–222. Stone, W. and Mathews, G. M. (1913). A list of the species of Australian birds described by John Gould, with the location of the type-specimens. Austral Avian Record, 1, 129–80. Stones, A. J., Lucking, R. S., Davidson, P. J., and Raharjaningtrah, W. (1997). Checklist of the birds of the Sula Islands (1991), with particular reference to Taliabu Island. Kukila, 9, 37–55. Stonor, C. R. (1937). An historical note on the Menebiki Ground-cuckoo (Centropus menebiki). Ibis, Ser. 14, vol. 1, 409–10. Storer, R. W. (1989). Notes on Paraguayan birds. Occasional Papers of the Museum of Zoology, University of Michigan, 719, 1–21. Storr, G. M. (1967). List of Northern Territory birds. Western Australian Museum, Special Publication, 4. Storr, G. M. (1977). Birds of the Northern Territory. Western Australian Museum, Special Publication, 7. Storr, G. M. (1980). Birds of the Kimberley Division, Western Australia. Western Australian Museum, Special Publication, 11. Storr, G. M. (1984a). Birds of the Pilbara Region, Western Australia. Records of the Western Australian Museum, Supplement, 16. Storr, G. M. (1984b). Revised list of Queensland birds. Records of the Western Australian Museum, Supplement, 19. Storr, G. M. (1985). Birds of the Gascoyne Region, Western Australia. Records of the Western Australian Museum, Supplement, 21. Storr, G. M. (1991). Birds of the South-West Division of Western Australia. Records of the Western Australian Museum, Supplement, 35.

Storrs, H. T. (1944). Indian Cuckoo laying in the nest of the Himalayan Black Drongo. Journal of the Bengal Natural History Society 18, 116–20. Stotz, D. F., Bierregaard, R. O., Cohn-Haft, M., Petermann, P., Smith, J., Whittaker, A., and Wilson, S. V. (1992). The status of North American migrants in central Amazonian Brazil. Condor, 94, 608–21. Stotz, D. F., Fitzpatrick, J. W., Parker, T. A., and Moskovits, D. K. (1996). Neotropical birds, ecology and conservation. University of Chicago Press, Chicago. Stotz, D. E., Lanyon, S. M., Schulenberg, T. S., Willard, D. E., Peterson, A. T., and Fitzpatrick, J. W. (1997). An avifaunal survey of two tropical forest localities on the Middle Rio Jiparaná, Rondônia, Brazil. Ornithological Monographs, 48, 763–81. Strahan, R., ed. (1994). Cuckoos, nightbirds and kingfishers of Australia. Angus and Robertson, Sydney. Strange, M. (2000). A photographic guide to the birds of Southeast Asia. Periplus, Singapore. Strange, M. (2001). A photographic guide to the birds of Indonesia. Periplus, Singapore. Strange, M. and Jeyarajasingam, A. (1993). Birds, a photographic guide to the birds of peninsular Malaysia and Singapore. Sun Tree, Singapore. Strassmann, J. E., Queller, D. C., Solis, C. R., and Hughes, C. R. (1991). Relatedness and queen number in the Neotropical wasp, Parachartergus colobopterus. Animal Behaviour, 42, 461–70. Strassmann, J. E., Gastreich, K. R., Queller, D. C., and Hughes, C. R. (1992). Demographic and genetic evidence for cyclical changes in queen number in a Neotropical wasp Polybia emaciata. American Naturalist, 140, 363–72. Straube, F. C. and Bornschein, M. R. (1989). A contribuição de André Mayer à História Natural no Paraná. I. Sobre uma coleção de aves do extremo noroeste do Paraná e sud do Mato Grosso do Sul. Arquivos de Biologia e Tecnologia, 32, 441–71. Straube, F. C. and Bornschein, M. R. (1995). New or noteworthy records of birds from northwestern Paraná and adjacent areas (Brazil). Bulletin of the British Ornithologists’ Club, 115, 219–25. Strazds, M. (1996). Latvijas Meza Putni. Latvias Ornitologijas Biedriba, Riga.

Bibliography 597 Stresemann, E. (1912). Ornithologische Miszellen aus dem Indo-Australischen Gebiet. x. Die Formen von Cacomantis merulinus und Cacomantis sepulcralis, xi. Das geographische Variieren von Centropus bengalensis (Gm.). Novitates Zoologicae, 19, 332–5, 336–9.

Stresemann, E. (1940). Die Vögel von Celebes. Teil IIIB. Systematik und Biologie. Journal für Ornithologie , 88, 389–487.

Stresemann, E. (1913a). Ornithologische Miszellen aus dem Indo-Australischen Gebiet. Novitates Zoologicae, 20, 289–324.

Stresemann, E. (1975). Ornithology from Aristotle to the present, (translated by H. J. and C. Epstein, ed. by G. W. Cottrell, foreword and epilogue by E. Mayr). Harvard University Press, Cambridge, Massachusetts.

Stresemann, E. (1913b). Die Vögel von Bali. Aus den zoologischen Ergebnissen der II. Freiburger MolukkenExpedition. Novitates Zoologicae, 20, 325–87. Stresemann, E. (1914). Die Vögel von Seran (Ceram). Aus den zoologischen Ergebnissen der II. Freiburger Molukken-Expedition. Novitates Zoologicae, 21, 25–153.

Stresemann, E. (1954). Ferdinand Deppe’s travels in Mexico, 1824–1829. Condor, 56, 86–92.

Stresemann, E. and Heinrich, G. (1939). Die Vögel des Mount Victoria. Mitteilungen aus der Zoologischen Museum in Berlin, 24, 151–264. Stresemann, E. and Stresemann, V. (1966). Die Mauser der Vogel. Journal für Ornithologie , Sonderheft, 107.

Stresemann, E. (1924a). Mutationsstudien. VI. Cuculus clamosus Latham. Journal fürOrnithologie , 72, 77–8.

Stresemann, E. and Stresemann, V. (1969). Die Mauser der Schopfkuckucke (Clamator). Journal für Ornithologie , 110, 192–204.

Stresemann, E. (1924b). Mutationsstudien. VII. Cacomantis infuscatus (Hartlaub). Journal für Ornithologie , 72, 78–9.

Stresemann, V. and Stresemann, E. (1961). Die Handschwingen-Mauser der Kuckuck (Cuculidae). Journal für Ornithologie , 102, 317–52.

Stresemann, E. (1924c). Mutationsstudien. VIII. Clamator serratus (Sparrman). Journal für Ornithologie , 72, 79–83.

Struhsaker, T. T. (1997). Ecology of an African rain forest. University of Florida Press, Gainesville.

Stresemann, E. (1927a). Die Wirtsvögel von Eudynamis scolopacea salvadorii Hartert. Ornithologische Monatsberichte, 35, 86.

Styles, C. (1995). Notes on the bird species observed feeding on mopane worms. Birding in Southern Africa, 47, 53–4.

Stresemann, E. (1927b). Centropus phasianinus thierfelderi ssp. nov. Ornithologische Monatsberichte, 35, 111–112.

Subrahmanyam, M. V. V. and Krishnamoorthy, R. V. (1981). Physical characterisation of the song of the Koel Eudynamys scolopacea. Journal of the Bombay Natural History Society, 77, 247–52.

Stresemann, E. (1928). Cuculus canorus bangsi Oberholser im Winterquartier. Ornithologische Monatsberichte, 36, 19. Stresemann, E. (1930a). Neue Vogelrassen aus Kwangsi II. Ornithologische Monatsberichte, 38, 47–9. Stresemann, E. (1930b). Eine dritte Vogelsammlung aus Kwangsi. Journal für Ornithologie , 78, 301–8. Stresemann, E. (1931). Vorläufiges über die ornithologischen Ergebnisse der Expedition Heinrich 1930–1931. V. Zur Ornithologie von Halmahera und Batjan. Ornithologische Monatsberichte, 39, 167–71.

Supari, S. (2003). Birds songs of Singapore (CD). Nature Society, Singapore. Sutton, G. M. (1913). A pet Road-runner. Bird-Lore, 15, 324–6. Sutton, G. M. (1915). Suggestive methods of bird-study: pet Road-runners. Bird-Lore, 17, 57–61. Sutton, G. M. (1922). Notes on the Road-runner at Fort Worth, Texas. Wilson Bulletin, 34, 3–20. Sutton, G. M. (1951). Birds and an ant army in southern Tamaulipas. Condor, 53, 16–8.

Stresemann, E. (1934). Sauropsida: Aves. Handbuch der Zoologie. Band 7 (2). (eds. W. Kükenthal and T. Krumbach). Walter de Gruyter, Berlin and Leipzig.

Sutton, G. M. (1967). Oklahoma birds. University of Oklahoma Press, Norman.

Stresemann, E. (1939). Zur Systematik der Gattung Centropus (Mutationsstudien XXVIII). Journal für Ornithologie, 87, 61–4.

Sutton, G. M. (1973). Winter food of a central Oklahoma Roadrunner. Bulletin of the Oklahoma Ornithological Society, 5, 30.

598 Bibliography Sutton, G. M., Lea, R. B., and Edwards, E. P. (1950). Notes on the ranges and breeding habits of certain Mexican birds. Bird-Banding, 21, 45–59.

of a cuckoo-host association and the evolution of host defenses. American Naturalist, 142, 819–39.

Sutton, R. W. W. (1965). Notes on Ghanaian birds seen in 1964. Ibis, 107, 252–3.

Taplin, A. and Beurteaux, Y. (1992). Aspects of the breeding biology of the Pheasant Coucal Centropus phasianinus. Emu, 92, 141–6.

Swainson, W. (1838). Animals in menageries. Longman, London.

Tarboton, W. (1975). African Cuckoo parasitising Forktailed Drongo. Ostrich, 46, 186–8.

Swarth, H. S. (1931). The avifauna of the Galapagos Islands. California Academy of Sciences, San Francisco.

Tarboton, W. (1986). African Cuckoo: the agony and ecstasy of being a parasite. Bokmakierie, 38, 109–11.

Swierczewski, E. V. and Raikow, R. J. (1981). Hind limb morphology, phylogeny, and classification of the Piciformes. Auk, 98, 466–80. Swinhoe, R. (1860). The ornithology of Amoy. Ibis, 2, 130–3. Swinhoe, R. (1863). The ornithology of Formosa, or Taiwan. Ibis, 5, 250–311, 377–434. Swinhoe, R. (1870). On four new species of birds from China. The Annals and Magazine of Natural History, 4th Ser., vol. 6, 152–4. Swofford, D. L. (2001). PAUP*. Phylogenetic Analysis Using Parsimony (*and other methods). Version 4. Sinauer Associates, Sunderland, Massachusetts. Swofford, D. L., Olsen, G. J., Waddell, P. J., and Hillis, D. M. (1996). Phylogenetic inference. In Molecular systematics, (eds. D.M. Hillis, C. Moritz, and B. K. Mable), pp. 407–514. Sinauer, Sunderland, Massachusetts. Swynnerton, C. F. M. (1916). On the colouration of the mouths and eggs of birds. I. The mouths of birds. II. On the colouration of eggs. Ibis, Ser. 10, vol. 4, 264–94, 529–606. Swynnerton, C. F. M. (1918). Rejection by birds of eggs unlike their own; with remarks on some of the cuckoo problems. Ibis, Ser. 10, vol. 6, 27–54. Symons, G. (1962). Indian Mynah host to cuckoo. African Wild Life, 16, 343. Symons, G. (1978). Red-chested Cuckoos and Cape Robins. Natal Bird Club Newsletter, 251, 2–3. Takasu, F. (1998). Why do all host species not show defense against avian brood parasitism: evolutionary lag or equilibrium? American Naturalist, 151, 193–205. Takasu, F., Kawasaki, K., Nakamura, H., Cohen, J. E., and Shigesada, N. (1993). Modeling the population dynamics

Tarboton, W. (2001). A guide to the nests and eggs of southern African birds. Struik, Cape Town. Tarboton, W. R., Kemp, M. I., and Kemp, A. C. (1987). Birds of the Transvaal. Transvaal Museum, Pretoria. Tarburton, M. K. (1993). Is courtship a necessary prelude to mating? Australian Bird Watcher, 15, 96–7. Taylor, P. M. (1990). The folk biology of the Tobelo people: a study in folk classification. Smithsonian Contributions to Anthropology, 34. ten Cate, C. and Vos, D. R. (1999). Sexual imprinting and evolutionary processes in birds: a re-assessment. Advances in the Study of Animal Behaviour, 28, 1–31. Terborgh, J. W., Faaborg, J., and Brockmann, H. J. (1978). Island colonization by Lesser Antillean birds. Auk, 95, 59–72. Terborgh, J. W., Fitzpatrick, J. W., and Emmons, L. (1984). Annotated checklist of bird and mammal species of Cocha Cashu Biological Station, Manu National Park, Peru. Fieldiana Zoology, New Series, 21. Terborgh, J. W., Robinson, S. K., Parker, T. A., Munn, C. A., and Pierpont, N. (1990). Structure and organization of an Amazonian forest bird community. Ecological Monographs, 60, 213–38. Teuschl, T., Taborsky, B., and Taborsky, M. (1998). How do cuckoos find their hosts? the role of habitat imprinting. Animal Behaviour, 56, 1425–33. Tewksbury, J. J. and Nabhan, G. P. (2001). Directed deterrence by capsaicin in chillies. Nature, 412, 403–4. Thewlis, R. M., Duckworth, J. W., Anderson, G. Q. A., Dvorak, M., Evans, T. D., Nemeth, E., Timmins, R. J., and Wilkinson, R. J. (1996). Ornithological records from Laos, 1992–1993. Forktail, 11, 47–100. Thibault, J. C. and Rives, C. (1975). Birds of Tahiti. Translated by D. T. Holyoak. Les Éditions du Pacifique, Papeete, Tahiti.

Bibliography 599 Thiede, W. (1994). Kuckucks-Fragen. Teil 7. Der Dreisilbenkuckuck. Ornithologische Mitteilungen, 46, 102–4. Thiollay, J.-M. (1985). The birds of Ivory Coast. Malimbus, 7, 1–59. Thiollay, J.-M. and Jullien, M. (1998). Flocking behaviour of foraging birds in a Neotropical rain forest and the antipredator defence hypothesis. Ibis, 150, 382–94. Thomas, B. T. (1978). The Dwarf Cuckoo in Venezuela. Condor, 80, 105–6. Thomas, B. T. (1979). The birds of a ranch in the Venezuelan llanos. In Vertebrate ecology in the northern Neotropics, (ed. J. F. Eisenberg), pp. 213–32. Smithsonian Institution Press, Washington, D. C. Thomas, B. T. (1995). Black-billed Cuckoo parasitizes the nest of a Yellow-breasted Chat. Raven, 66, 3–5. Thomas, J. (1991). Birds of the Korup National Park, Cameroon. Malimbus, 13, 11–23. Thomas, R. and Thomas, S. (1994). Birds of Sulawesi, Lesser Sundas and Sabah. privately published cassette, U.K. Thompson, J. D., Higgins, D. G., and Gibson, T. J. (1994). Clustal W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, positions-specific gap penalties and weight matrix choice. Nucleic Acids Research, 22, 4673–80. Thompson, M. C. (1966). Birds from North Borneo. University of Kansas Publications Museum of Natural History, 17, 377–433. Thompson, P. M., Harvey, W. G, Johnson, D. L., Millin, D. J., Rashid, S. M. A., Scott, D. A., Stanford, C., and Woolner, J. D. (1993). Recent notable bird records from Bangladesh. Forktail, 9, 13–44. Thorington, R. W. and Thorington, E. M. (1989). Postcranial proportions of Microsciurus and Sciurillus, the American pygmy tree squirrels. In Advances in Neotropical mammalogy, (eds. J. F. Eisenberg and K. H. Redford), pp. 125–36. Sandhill Crane Press, Gainesville, Florida.

Tikader, B. K. (1984). Birds of Andaman and Nicobar Islands. Zoological Survey of India, Calcutta. Tojo, H., Nakamura, S., and Higuchi, H. (2002). Gape patches in Oriental Cuckoo Cuculus saturatus. Ornithological Science, 1, 145–9. Tolhurst, L. P. (1992). Extension to the range of the Lesser Black Coucal, Centropus bernsteini. Muruk, 5, 92–3. Tomek, T. (1985). Materials for the breeding avifauna of the Democratic People’s Republic of Korea. Results of Expedition ’83. Acta Zoologica Cracov, 29, 187–218. Topp, A. (1999). Photonews: Oriental Cuckoos Cuculus saturatus. Alula, 5, 90–1. Tostain, O., Dujardin, J.-L., Erard, C., and Thiollay, J.-M. (1992). Oiseaux de Guyane. Société d’Études Ornithologiques, Brunoy, France. Trainer, J. M. and Will, T. C. (1984). Avian methods of feeding on Bursera simaruba (Burseraceae) fruits in Panama. Auk, 101, 193–5. Trautman, M. B. (1940). The birds of Buckeye Lake, Ohio. Miscellaneous Publications of the Museum of Zoology, University of Michigan, 44. Traylor, M. A. (1958). Birds of northeastern Peru. Fieldiana: Zoology, 35, 87–141. Traylor, M. A. (1960). Notes on the birds of Angola, non-passeres. Publicações Culturais da Companhia de Diamantes de Angola, Lisboa, 54, 129–86. Traylor, M. A. (1967). A collection of birds from Szechwan. Fieldiana: Zoology, 53, 1–67. Tree, A. J. (1983). Another record of nominate Jacobin Cuckoo from Zimbabwe. Honeyguide, 114–115 , 59. Trevor, S. and Lack, P. (1976). Great Spotted Cuckoo parasitizing Superb Starling. East African Natural History Society Bulletin, 1976, 50. Turner, J. R. G. (1984). Mimicry: the palatability spectrum and its consequences. Symposium of the Royal Entomological Society of London, 11, 141–61.

Thornton, J. (1996). Krakatau. Harvard University Press, Cambridge, Massachusetts.

Turner, J. R. G. (1995). Mimicry as a model for coevolution. In Biodiversity and evolution (eds. R. Arai, M. Kato, and Y. Doi), pp. 13l–50. National Science Museum. Foundation, Tokyo.

Thurber, W. A., Serrano, J. F., Sermeño, A., and Benitez, M. (1987). Status of uncommon or previously unreported birds of El Salvador. Proceedings of the Western Foundation of Vertebrate Zoology, 3, 109–293.

Tweit, R. C. and Tweit, J. C. (1986). Urban development effects on the abundance of some common resident birds of the Tucson area of Arizona. American Birds, 40, 431–6.

600 Bibliography Tyler, C. (1969). Avian egg-shells: their structure and characteristics. International Review of General and Experimental Zoology, 4, 81–130.

van Someren, V. G. L. (1925). What is the true relationship of Klaas’s Cuckoo to Chrysococcyx cupreus and Lampromorpha caprius? Ibis, Ser. 12, vol. 1, 660–2.

Tymstra, R., Connop, C., and Tshering, C. (1997). Some bird observations from central Bhutan, May 1994. Forktail, 12, 49–60.

van Someren, V. G. L. (1949). The birds of Bwamba. Uganda Journal, Supplement, 13, 1–111.

Urano, E., Yamagishi, S., Arimisa, A. A., and Arafara, S. A. (1994). Different habitat use among three sympatric species of couas Coua cristata, C. coquereli and C. ruficeps in western Madagascar. Ibis, 136, 485–7. Urban, E. K. and Brown, L. H. (1971). A checklist of the birds of Ethiopia. Department of Biology, Haile Sellassie I University, Addis Ababa. Valverde, J. A. (1971). Notas sobre la biologia de reproduccion del crialo Clamator glandarius (L.). Ardeola, no. especial, 591–647. van Balen, S. (1997). Faunistic notes from Kayan Mentarang with new records for Kalimantan. Kukila, 9, 108–13. van Balen, S. and Prentice, C. (1997). Birds of the Negara River Basin, South Kalimantan, Indonesia. Kukila, 9, 81–107. van Bemmel, A. C. V. (1948). A faunal list of the birds of the Moluccan Islands. Treubia, 19, 323–402. van Bemmel, A. C. V. and Voous, K. H. (1951). On the birds of the islands of Muna and Buton, S. E. Celebes. Treubia, 21, 27–104. van den Berg, A. B. (1994). WP reports. Dutch Birding, 16, 77–83. van den Berg, A. B. and Bosman, C. A. W. (1986). Supplementary notes on some birds of Lore Lindu Reserve, Central Sulawesi. Forktail, 1, 7–13. van Marle, J. G. and Voous, K. H. (1988). The birds of Sumatra, an annotated check-list. BOU Check-list No. 10, British Ornithologists’ Union, London.

van Someren, V. G. L. (1956). Days with birds. Fieldiana: Zoology, 38. van Someren, V. G. L. and van Someren, G. R. C. (1945). Evacuated weaver colonies and notes on the breeding ecology of Euodice cantans Gmelin and Amadina fasciata Gmelin. Ibis, 87, 33–44. van Tuinen, M. and S. B. Hedges (2001). Calibration of avian molecular clocks. Molecular Biology and Evolution, 18, 206–13. van Tuinen, M., Sibley, C. G., and Hedges, S. B. (2000). The early history of modern birds inferred from DNA sequences of nuclear and mitochondrial ribosomal genes. Molecular Biology and Evolution, 17, 451–7. Van Tyne, J. and Sutton, G. M. (1937). The birds of Brewster County, Texas. Miscellaneous Publications of the Museum of Zoology University of Michigan, 37. Varga, F. (1994). Cuckoo observations around the source of River Zagyva. Fernec Varga, Salgótarján, Hungary. Vasamies, H. (1998). The first Oriental Cuckoos in Western Europe. Alula, 3, 114–6. Vaurie, C. (1949). A revision of the bird family Dicruridae. Bulletin of the American Museum of Natural History, 93, 119–342. Vaurie, C. (1965). The birds of the Palearctic fauna, vol. 2, Non Passeriformes. H. F. and G. Witherby, London. Vehrencamp, S. L. (1976). The evolution of communal nesting in Groove-billed Anis. PhD dissertation, Cornell University. Vehrencamp, S. L. (1977). Relative fecundity and parental effort in communally nesting anis, Crotophaga sulcirostris. Science, 197, 403–5.

van Rossem, A. J. (1930). Some geographic variations in Piaya cayana. Transactions of the San Diego Society of Natural History, 6, 209–10.

Vehrencamp, S. L. (1978). The adaptive significance of communal nesting in Groove-billed Anis (Crotophaga sulcirostris). Behavioral Ecology and Sociobiology, 4, 1–33.

van Rossem, A. J. (1934). Critical notes on Middle American birds. Bulletin of the Museum of Comparative Zoology, Harvard, 77, 387–490.

Vehrencamp, S. L. (1982a). Body temperatures of incubating versus non-incubating Roadrunners. Condor, 84, 203–7.

van Rossem, A. J. (1938). The Groove-billed Ani of Lower California and northwestern Mexico. Condor, 40, 91.

Vehrencamp, S. L. (1982b). Testicular regression in relation to incubation effort in a tropical cuckoo. Hormones and Behavior, 16, 113–20.

Bibliography 601 Vehrencamp, S. L., Stiles, G. F., and Bradbury, J. W. (1977). Observations on the foraging behavior and avian prey of the Neotropical carnivorous bat, Vampyrum spectrum. Journal of Mammalogy, 58, 469–78. Vehrencamp, S. L., Bowen, B. S., and Koford, R. R. (1986). Breeding roles and pairing patterns within communal groups of Groove-billed Anis. Animal Behaviour, 34, 347–66. Vehrencamp, S. L., Koford, R. R., and Bowen, B. S. (1988). The effect of breeding-unit size on fitness components in Groove-billed Anis. In Reproductive success, (ed. T. H. Clutton-Brock), pp. 291–304. University of Chicago Press, Chicago. Verheijen, J. A. J. (1964). Breeding season on the island of Flores, Indonesia. Ardea, 2, 194–201. Verheyen, R. (1953). Oiseaux. Exploration du Parc National de l’Upemba, 19. Institut des Parcs Nationaux du Congo Belge, Bruxelles. Verheyen, R. (1956). Contribution à l’anatomie et a la systématique des touracos (Musophagi) et des coucous (Cuculiformes). Bulletin Institut Royal des Sciences Naturelles de Belgique, 32 (23), 1–28. Verhoeye, J. and Holmes, D. A. (1998). The birds of the islands of Flores - a review. Kukila, 10, 3–59. Vernon, C. J. (1967). The acacia savanna avifauna of Bulawayo, Rhodesia. South African Avifauna Series, 43. Vernon, C. J. (1970a). Pre-incubation embryonic development and egg “dumping” by the Jacobin Cuckoo. Ostrich, 41, 259–60. Vernon, C. J. (1970b). New host species for three cuckoos. Ostrich, 41, 258. Vernon, C. J. (1971). Notes on the biology of the Black Coucal. Ostrich, 42, 242–58. Vernon, C. J. (1982). Notes on the breeding of the Striped Crested Cuckoo. Honeyguide, 111–112 , 10–1. Vernon, C. J. (1984). The breeding biology of the Thickbilled Cuckoo. Proceedings of the V Pan-African Ornithological Congress, 825–40. Vernon, C. J., Macdonald, I. A. W., and Dean, W. R. J. (1990). The birds of the Haroni-Lusitu. Honeyguide, 36, 14–35. Vernon, C. J., Herremans, M., Underhill, L. G., and Berruti, A. (1997). Cuculidae: cuckoos and coucal. In The atlas of southern African birds, vol. 1. Non-passerines, (eds. J. A. Harrison, D. G. Allan, L. G. Underhill,

M. Herremans, M., A. J. Tree, V. Parker and C. J. Brown), pp. 544–73. BirdLife South Africa, Johannesburg. Vernon, D. P. (1968). Birds of Brisbane and environs. Queensland Museum, Brisbane. Viksnes, J. (1989). Latvian breeding bird atlas 1980–1984 . Zinatne, Riga. Vincent, A. W. (1946). On the breeding habits of some African birds. Ibis, 88, 48–67. Vincent, F. (1965). Le coucou emeraude (Chrysococcyx cupreus intermedius Hartlaub 1857) parasite d’un nid de souimanga (Nectariniidae) au Congo (Brazzaville). L’Oiseau et la Revue Française d’Ornithologie, 35, 81–6. Vincent, J. (1934). The birds of northern Portuguese East Africa. Comprising a list of, and observations on, the collections made during the British Museum Expedition of 1931–1932, part V. Ibis, Ser. 13, vol. 4, 757–99. Violani, C., Zanazzo, G., and Pandolfi, M. (1997). La collezione ornitologica di Tommaso Salvadori. Servizio Beni e Attività Culturali, Comune di Fermo, Italy. de Visscher, M. N. and Moratorio, M. (1984). Notes on the feeding behaviour of the Smooth-billed Ani, Crotophaga ani. Gerfaut, 74, 71–4. Vogel, C. J., Sweet, P. R., Le Manh Hung, and Hurley, M. M. (2003). Ornithological records from Ha Giang province, north-east Vietnam, during March–June 2000. Forktail, 19, 21–30. Voipio, P. (1953). The hepaticus variety and the juvenile types of the Cuckoo. Ornis Fennica, 30, 97–117. Voisin, C. and Voisin, J.-F. (1991). Un spécimen méconnu de Coua de Delalande Coua delalandei Temminck. L’Oiseau et la Revue Française d’Ornithologie, 61, 341–2. von Berlepsch, H. (1901). Systematisches Verzeichnis der von Herrn Professor Willy Kükenthal währund seiner Reisen im Malayischen Archipel im Jahre 1894 auf den nördlichen Molukken-Inseln gesammelten Vogelbälge. Abhandlungen von der Senckenbergischen Naturforschenden Gesellschaft, 25, 299–313. von Frisch, O. (1969). Die Entwicklung des Häherkuckucks (Clamator glandarius) im Nest der Wirtsvögel und seine Nachtzuch in Gefangenschaft. Zeitschrift für Tierpsychologie , 26, 641–50. von Frisch, O. (1973). Ablenkungsmanöver bei der Eiablage des Häherkuckucks (Clamator glandarius). Journal für Ornithologie , 114, 129–31.

602 Bibliography von Frisch, O. and von Frisch, H. (1967). Beobachtungen zur Brutbiologie und Jugendentwicklung des Häherkuckucks (Clamator glandarius). Zeitschrift für Tierpsychologie , 24, 129–36. von Ihering, H. (1914). Biologie der brasilianischen Cuculidae. Rivista do Museum Paulista, 9, 391–410. Voous, K. H. (1951). The egg of Cacomantis sonneratii schlegeli (Junge). Ardea, 39, 237–8. Voous, K. H. (1957). The birds of Aruba, Curaçao, and Bonaire. Studies on the Fauna of Curaçao and other Caribbean Islands, 7, 1–259. Voous, K. H. (1983). Birds of the Netherlands Antilles. (2nd edn.) De Walburg Pers, Utrecht. Vorderman, A. G. (1898). Molukken-Vogels. Nederlands Tijdschrift voor Natuurkunde, Nederlandsch Indië, 58, 169–252. Vuilleumier, F. (1993). Notes on birds observed in beech (Fagus) forests in the Maoershan Natural Reserve, Guangxi Autonomous Region, China. Bulletin of the British Ornithologists’ Club, 113, 152–66. Wait, W. E. (1925). Manual of the birds of Ceylon. Ceylon Journal of Science, Colombo. Wakeley, J. (1996). The excess of transitions among nucleotide substitutions: new methods of estimating transition bias underscore its significance. Trends in Ecology and Evolution, 11, 158–63. Walden, A. (1869). On the Cuculidae described by Linnaeus and Gmelin, with a sketch of the genus Eudynamis. Ibis, 1869, 324–46. Waldrigues, A. and Ferrari, I. (1979). Estudo cromossômico em Crotophaga ani (Aves). Ciência e Cultura, 31, 03–G.1.6. Waldrigues, A. and Ferrari, I. (1980). Estudo cromossômico em Guira guira (Aves). Ciência e Cultura, 32, 267–G.1.6.

Wallace, A. R. (1869). The Malay archipelago. Macmillan, London. Wallace, A. R. (1889). Darwinism; an exposition of the theory of natural selection with some its applications. Macmillan, London. Wallace, G. E., Gonzales Alonso, H., McNicholl, M. K., Rogriguez Batista, D., Oviedo Prieto, R., Llanes Sosa, A., Sanchez Oria, B., and Wallace, E. A. H. (1996). Winter surveys of forest-dwelling Neotropical migrant and resident birds in three regions of Cuba. Condor, 98, 745–68. Walters, M. (1996). The eggs of Carpococcyx renauldi: a correction. Bulletin of the British Ornithologists’ Club, 116, 271. Walters, M. (2003). A concise history of ornithology. Yale University Press, New Haven. Wardill, J. C., Fox, P. S., Hoare, D. J., Marthy, W., and Anggraini, K. (1998). Birds of the Rawa Aopa Watumokai National Park, south-east Sulawesi. Kukila, 10, 91–114. Warner, D. W. (1951) A new race of the cuckoo, Chalcites lucidus, from the New Hebrides islands. Auk, 68, 106–7. Warren, J. W. (1960). Temperature fluctuations in the Smooth-billed Ani. Condor, 62, 293–4. Warren, R. L. M. (1966). Type-specimens of birds in the British Museum (Natural History). Vol. 1, Non-passerines. British Museum (Natural History), London. Watling, D. (1983). Ornithological notes from Sulawesi. Emu, 83, 247–61. Watts, R. A. (1993). Male Pallid Cuckoo feeding juvenile cuckoo. Australian Birds, 26, 101–2. Wauer, R. H. (1996). A birder’s West Indies. University of Texas Press, Austin.

Waldrigues, A. and Ferrari, I. (1982). Karyotypic study of cuculiform birds. I. Karyotype of the Smooth-billed Ani (Crotophaga ani). Revista Brasileira de Genetica, 5, 121–9.

Wauer, R. H. and Ligon, J. D. (1977). Distributional relations of breeding avifauna of four southwestern mountain ranges. Symposium on the Biological Resources of the Chihuahuan desert Region, National Park Service, 3, 567–78.

Waldrigues, A., Ferrari, I., and Faustino Neto, A. (1983). Estudo cariotípico em duas espécies de Cuculiformes americanos (Aves). Acta Amazonica, 13, 37–50.

Weathers, W. W. (1983). Birds of southern California’s Deep Canyon. University of California Press, Berkeley.

Wallace, A. R. (1853). A narrative of travels on the Amazon and Rio Negro. Reeve, London.

Webster, M. and Fook, Chew Yen. (1997). A photographic guide to birds of Thailand. New Holland Publishers, London.

Wallace, A. R. (1862). List of birds from the Sula Islands (east of Celebes), with descriptions of the new species. Proceedings of the Zoological Society of London, 1862, 333–46.

Weigel, R. D. (1963). Oligocene birds from Saskatchewan. Quarterly Journal of the Florida Academy of Sciences, 26, 257–62.

Bibliography 603 Weins, J. J. and Hollingsworth, B. D. (2000). War of the iguanas: conflicting molecular and morphological phylogenies and long branch attraction in iguanid lizards. Systematic Biology, 49, 143–59. Weisler, M. I. (1995). Henderson Island prehistory: colonization and extinction on a remote Polynesian island. Biological Journal of the Linnean Society, 56, 377–404. Wells, D. R. (1972). The genus Cuculus: two amendments to the ‘Handbook of the birds of India and Pakistan.’ Journal of the Bombay Natural History Society, 69, 179–85.

Survey of Porto Rico and the Virgin Islands, 9 (4), 409–571. Wetmore, A. (1968). Birds of the Republic of Panama. Part 2. Columbidae (Pigeons) to Picidae (Woodpeckers). Smithsonian Institution Press, Washington, D.C. Wetmore, A. and Swales, B. H. (1931). The birds of Haiti and the Dominican Republic. United States National Museum Bulletin 155. Wheeler, W. C. (1996). Sequence alignment, parameter sensitivity and the phylogenetic analysis of molecular data. Systematic Biology, 44, 321–31.

Wells, D. R. (1976). Some bird communities in western Sabah, with distributional records, March 1975. Sarawak Museum Journal, 24, 277–86.

Whistler, H. (1926). The birds of the Kangra District, Punjab. Part 2. Ibis, 12th Ser. vol. 2, 724–83.

Wells, D. R. (1982). A confirmation of the specific relations of Cuculus saturatus insulindae Hartert. Bulletin of the British Ornithologists’ Club, 102, 62–3.

Whistler, H. (1928). The migration of the Pied Crested Cuckoo (Clamator jacobinus). Journal of the Bombay Natural History Society, 33, 136–45.

Wells, D. R. (1999). The birds of the Thai-Malay peninsula. vol. 1, Non-passerines. Academic Press, New York.

Whistler, H. (1944). The avifaunal survey of Ceylon. Spolia Zeylanica, 23, 119–321.

Wells, D. R. and Becking, J. H. (1975). Vocalizations and status of Little and Himalayan Cuckoos, Cuculus poliocephalus and C. saturatus, in southeast Asia. Ibis, 117, 366–71.

Whistler, H. and Kinnear, N. B. (1934). The Vernay Scientific Survey of the Eastern Ghats (Ornithological section), part 9. Journal of the Bombay Natural History Society, 37, 515–28.

Welty, J. C. and Baptista, L. (1988). The life of birds. (4th edn.) Saunders, New York.

White, C. M. N. (1977). Notes on some non-passerine birds of Wallacea. Bulletin of the British Ornithologists’ Club, 97, 99–103.

Wendelken, P. W. and Martin, R. F. (1986). Recent data on the distribution of birds in Guatemala. Bulletin of the British Ornithologists’ Club, 106, 16–21. Werth, I. (1947). The growth of a young Cuckoo. British Birds, 40, 331–4. West, O., Wright, F. B., and Symons, G. (1964). The birds of Weenen County, Natal. South African Avifauna Series, 14. Westneat, D. F. and Sargent, R. C. (1996). Sex and parenting: the effects of sexual conflict and parentage on parental strategies. Trends in Ecology and Evolution, 11, 87–91. Westneat, D. F. and Sherman, P. W. (1993). Parentage and the evolution of parental behavior. Behavioral Ecology, 4, 68–77.

White, C. M. N. and Bruce, M. D. (1986). The birds of Wallacea (Sulawesi, the Moluccas and Lesser Sunda Islands, Indonesia). BOU Check-list No. 7, British Ornithologists’ Union, London. White, T. (1984). A field guide to the bird songs of Southeast Asia, 2 tape cassettes. British Library, National Sound Archive. Whitson, M. A. (1971). Field and laboratory investigations of the ethology of courtship and copulation in the Greater Roadrunner (Geococcyx californianus—Aves, Cuculidae). PhD Dissertation, University of Oklahoma, Norman. Whitson, M. A. (1975). Courtship behavior of the Greater Roadrunner. Living Bird, 14, 215–55.

Wetmore, A. (1916). Birds of Porto Rico. United States Department of Agriculture Bulletin 326.

Whitson, M. A. (1983). The Roadrunner—clown of the desert. National Geographic, 163, 694–702.

Wetmore, A. (1926). Observations on the birds of Argentina, Paraguay, Uruguay, and Chile. United States National Museum Bulletin, 133.

Whittaker, A. (1995). First report of Coccyzus pumilus for Brazil (Cuculiformes: Cuculidae). Ararajuba, 3, 81.

Wetmore, A. (1927). The birds of Porto Rico and the Virgin Islands. New York Academy of Sciences, Scientific

Whittaker, A. and Oren, D. C. (1999). Important ornithological records from the Rio Juruá, western Amazonia, including twelve additions to the Brazilian

604 Bibliography avifauna. Bulletin of the British Ornithologists’ Club, 119, 235–60.

notes on the mammals. ICBP Study Report, no. 48, Cambridge, U. K.

Whitten, A. J., Mustafa, M., and Henderson, G. S. (1987). The ecology of Sulawesi. Gadjah Mada University Press, Yogyakarta, Indonesia.

Williams, G. C. (1966). Natural selection, the costs of reproduction, and a refinement of Lack’s principle. American Naturalist, 100, 687–90.

Whitten, T., Soeriaatmadja, R. E., and Afiff, S. A. (1996). The ecology of Java and Bali. Periplus Editions, Jakarta.

Williams, G. R. (1960). The birds of the Pitcairn Islands, Central South Pacific Ocean. Ibis, 102, 58–70.

Whittingham, L. A. and Dunn, P. O. (1998). Male parental effort and paternity in a variable mating system. Animal Behaviour, 55, 629–40.

Williams, T. (1995). The penguins. Bird Families of the World. Oxford University Press, Oxford.

Whittingham, L. A., Dunn, P. O., and Magrath, R. D. (1997). Relatedness, polyandry and extra-group mating in the cooperatively-breeding White-browed Scrubwren (Sericornis frontalis). Behavioral Ecology and Sociobiology, 40, 261–70. Wickler, S. and Spriggs, M. (1988). Pleistocene human occupation of the Solomon Islands, Melanesia. Antiquity, 62, 703–6. Wiens, J. A. (1965). Nest parasitism of the Dickcissel by the Yellow-billed Cuckoo in Marshall County, Oklahoma. Southwestern Naturalist, 10, 142. Wijesinghe, M. (1999). Nesting of Green-billed Coucals Centropus chlororhynchos in Sinharaja, Sri Lanka. Forktail, 15, 43–5. Wikelski, M., Hau, M., and Wingfield, J. C. (2000). Seasonality of reproduction in a Neotropical rain forest bird. Ecology, 81, 2458–72.

Willis, E. O. (1973). Survival rates for visited and unvisited nests of Bicolored Antbirds. Auk, 90, 263–7. Willis, E. O. (1974). Populations and local extinctions of birds on Barro Colorado Island, Panama. Ecological Monographs, 44, 153–69. Willis, E. O. (1980). Ecological roles of migratory and resident birds on Barro Colorado Island, Panama. In Migrant birds in the Neotropics: ecology, behavior, distribution, and conservation, (eds. A. Keast and E. S. Morton), pp. 205–25. Smithsonian Institution Press, Washington, D. C. Willis, E. O. (1982). Ground-cuckoos (Aves, Cuculidae) as army ant followers. Revista Brasileira de Biologia, 42, 753–6. Willis, E. O. (1983a). Piaya cuckoos (Aves, Cuculidae) as army ant followers. Revista Brasileira de Biologia, 43, 29–32. Willis, E. O. (1983b). Anis (Aves, Cuculidae) as army ant followers. Revista Brasileira de Biologia, 43, 33–6.

Wild Bird Society of Japan. (1980). The breeding bird survey in Japan 1978. Wild Bird Society of Japan, Tokyo.

Willis, E. O. and Eisenmann, E. (1979). A revised list of birds of Barro Colorado Island, Panama. Smithsonian Contributions in Zoology, 291.

Wild Bird Society of Japan. (1982). A field guide to the birds of Japan. Wild Bird Society of Japan, Tokyo.

Willis, E. O. and Oniki, Y. (1978). Birds and army ants. Annual Review of Ecology and Systematics, 9, 243–63.

Wildash, P. (1968). Birds of South Vietnam. Charles E. Tuttle, Rutland, Vermont.

Willis, E. O. and Oniki, Y. (1985). Bird specimens new for the state of São Paulo, Brazil. Revista Brasileira de Biologia, 45, 105–8.

Wiley, E. O., Siegel-Causey, D., Brooks, D. R., and Funk, V. A. (1991). The compleat cladist, a primer of phylogenetic procedures. University of Kansas Museum of Natural History Special Publication, 19. Wiley, R. H. and Rabenold, K. N. (1984). The evolution of cooperative breeding by delayed reciprocity and queuing for favorable social positions. Evolution, 38, 609–21. Wilkinson, A. S. (1947). A cuckoo in the nest. New Zealand Bird Notes, 2, 77–9. Wilkinson, R., Dutson, G., and Sheldon, R. (1991). The avifauna of Barito Ulu, Central Borneo, with additional

Willis, E. O. and Oniki, Y. (1990). Coccyzus euleri (Cabanis, 1873 (Aves, Cuculiformes)): proposed conservation of the specific name. Bulletin of Zoological Nomenclature, 47, 195–7. Willis, E. O. and Oniki, Y. (1991). Avifaunal transects across the open zones of northern Minas Gerais, Brazil. Ararajuba, 2, 41–58. Willis, E. O. and Oniki, Y. (1992). Losses of São Paulo birds are worse in the interior than in Atlantic forests. Ciência e Cultura, 44, 326–8.

Bibliography 605 Willis, E. O. and Oniki, Y. (1999). Nest building and early incubation in Squirrel Cuckoos (Piaya cayana). Ararajuba, 7, 23–5. Wilson, E. O. (1993). The diversity of life. W. W. Norton, New York. Wilson, R. G. (1992). Parasitism of Yellow-olive Flycatcher by the Pheasant Cuckoo. Euphonia, 1, 34–6.

Wragg, G. M. (1995). The fossil birds of Henderson Island, Pitcairn Group: natural turnover and human impact, a synopsis. In The Pitcairn Islands: biogeography, ecology and prehistory, (eds. T. G. Benton and T. Spence). Biological Journal of the Linnean Society, 56, 405–14.

Winterbottom, J. M. (1938). Further notes on some Northern Rhodesian birds. Ibis, 14th Ser., vol. 2, 269–77.

Wright, J., Parker, P. G., and Lundy, K. J. (1999). Relatedness and chick-feeding effort in the cooperatively breeding Arabian babbler. Animal Behaviour, 58, 779–85.

Winterbottom, J. M. (1962). Some manuscript notes of S. F. Townsend. Ostrich, 33, 66–71.

Wright, S. J. and Cornejo, F. H. (1990). Seasonal drought and leaf fall in a tropical forest. Ecology, 71, 1165–75.

Witherby, H. F., Jourdain F. C. R., Ticehurst, N. F., and Tucker, B. W. (1938). The handbook of British birds, vol. 2. Witherby, London.

Wunderle, J. M., Jr. (1981). Avian predation upon Anolis lizards on Grenada, West Indies. Herpetologia, 37, 104–8.

Witter, J. A. and Kulman, H. M. (1972). A review of the parasites and predators of tent caterpillars (Malacosoma spp.) in North America. University of Minnesota Agricultural Experimental Station Technical Bulletin, 289.

Wunderle, J. M., Jr. (1995). Responses of bird populations in a Puerto Rican forest to Hurricane Hugo: the first 18 months. Condor, 97, 879–96.

Wolfe, D. H. (1994). Yellow-billed Cuckoo hatched in Mourning Dove nest. Bulletin of the Oklahoma Ornithological Society, 27, 29–30. Wolfe, L. R. (1938). Birds of central Luzon. Auk, 55, 198–224. Wolters, H. E. (1975–1982). Die Vogelarten der Erde. Paul Parey, Hamburg.

Wylie, S. R. and Shelton, L. C. (1982). The Red-billed Ground Cuckoo Carpococcyx renauldi. AFA Watchbird, 9, 29–32. Wyllie, I. (1975). Study of Cuckoos and Reed Warblers. British Birds, 68, 369–78. Wyllie, I. (1981). The Cuckoo. B. T. Batsford, London.

Wood, N. A. (1951). The birds of Michigan. Miscellaneous Publications, Museum of Zoology, University of Michigan, 75.

Yamagishi, S. and Fujioka, M. (1986). Heavy brood parasitism by the Common Cuckoo Cuculus canorus on the Azure-winged Magpie Cyanopica cyana. Tori, Bulletin of the Ornithological Society of Japan, 34, 91–96.

Woodell, R. (1976a). Variation in juvenile plumage of Centropus toulou toulou (Muller) and Centropus toulou insularis Ridgway. Bulletin of the British Ornithologists’ Club, 96, 72–5.

Yamagishi, S., Honda, M., Eguchi, K., and Thorstrom, R. (2001). Extreme endemic radiation of the Malagasy vangas (Aves: Passeriformes). Journal of Molecular Evolution, 53, 39–46.

Woodell, R. (1976b). Notes on the Aldabran Coucals Centropus toulou insularis. Ibis, 118, 263–8.

Yamashina, Y. (1931). Die Vögel der Kurilen. Journal für Ornithologie, 79, 491–540.

Woods, R. S. (1960). Notes on the nesting of the Roadrunner. Condor, 62, 483–4.

Yamashina, Y. (1941). A natural history of Japanese birds, vol. 2. Iwanami, Tokyo.

Woodward, P. W. (1983). Behavioral ecology of fledgling Brown-headed Cowbirds and their hosts. Condor, 85, 151–63.

Yamashina, Y. (1946). The chromosomes of the Cuckoo, the Budgerigar, the Little Ringed Plover and the Chinese Bamboo Pheasant (In Japanese, English summary). Kromosomo, 1, 18–23.

Woolfenden, G. E. (1975). Florida Scrub Jay helpers at the nest. Auk, 92, 1–15. Worman, A. G. (1930). Male Emerald Cuckoo (Chrysococcyx cupreus intermedius) feeding young. Oological Record, 10, 76–7.

Yan, Z. (1995). Three new records of birds in Hunan Province. Chinese Journal of Zoology, 30, 46. Yang, Z. (1998). On the best evolutionary rate for phylogenetic analysis. Systematic Biology, 47, 125–33.

606 Bibliography Yen, K. Y. (1933). Les oiseaux du Kwangsi (Chine) (suite). L’Oiseau et la Revue Française d’Ornithologie, 3, 615–38. Young, C. G. (1929). A contribution to the ornithology of the coastland of British Guiana. Ibis, Ser. 12, vol. 5, 1–38. Yoon, M.-B. (2000). Wild birds of Korea. Kyo-Hak, Seoul. Yosef, R. (1997). First record of Great Spotted Cuckoo (Clamator glandarius) parasitizing Indian House Crow (Corvus splendens). Israel Journal of Zoology, 43, 397–9. Yosef, R. (2002). Second breeding record of the Great Spotted Cuckoo (Clamator glandarius) in Eilat. Sandgrouse 24, 142–44.

Natural History Publication 282, Zoology Series, 17, 233–480. Zimmer, K. J. and Hilty, S. L. (1997). Avifauna of a locality in the Upper Orinoco drainage of Amazonas, Venezuela. Ornithological Monographs, 48, 865–85. Zimmer, K. J., Parker, T. A., Isler, M. L., and Isler, P. R. (1997). Survey of a southern Amazonian avifauna: the Alta Floresta region, Mato Grosso, Brazil. Ornithological Monographs, 48, 887–918. Zimmerman, D. A. (1970). Roadrunner predation on passerine birds. Condor, 72, 473–6.

Yoshino, T. (1988). Kakko. [Cuckoos] Kara Shizenshiruzu 68. (Color Nature Series 68). Kaisei-Sha, Ichigaya, Tokyo. [in Japanese]

Zimmerman, D. A. (1972). The avifauna of the Kakamega Forest, western Kenya, including a bird population study. Bulletin of the American Museum of Natural History, 140, 259–339.

Yoshino, T. (1999). Kakko— Nihon no takurancho (Cuckoos— brood-parasitic birds in Japan ]. Bunichisogoshuppan, Tokyo. (in Japanese)

Zimmerman, D. A., Turner, D. A., and Pearson, D. J. (1996). Birds of Kenya and northern Tanzania. Princeton University Press, Princeton.

Zacharias, V. J. and Gaston, A. J. (1983). Breeding seasons of birds at Calicut, southwest India. Ibis, 125, 407–12.

Zink, R. M. (1996). Comparative phylogeography in North American birds. Evolution, 50, 308–17.

Zacharias, V. J. and Gaston, A. J. (1993). The birds of Wynaad, southern India. Forktail, 8, 11–23. Zarudny, N. A. (1914). Notiz über die Kuckucke von Turkestan (in Russian, with German summary). Messager Ornithologique (Ornithologische Mitteilungen), Moskva, 5, 105–15.

Zink, R. M., Blackwell-Rago, R. C., and Ronquist, F. (2000). The shifting roles of dispersal and vicariance in biogeography. Proceedings of the Royal Society of London, Series B, 267, 497–503. Zuñiga, J. M., Soler, M., and Camacho, I. (1983). Nota sobre nuevas especies parasitadas por el Críalo (Clamator glandarius) en España. Doñana, Acta Vertebratica , 10, 207–9.

Zeffer, A., Johansson, L. C., and Marmebro, A. (2003). Functional correlation between habitat use and leg morphology in birds (Aves). Biological Journal of the Linnean Society, 79, 461–84.

Zuñiga, J. M. and Redondo, T. (1992a). No evidence for variable duration of sympatry between the Great Spotted Cuckoo and its Magpie host. Nature, 359, 410–1.

Zetra, B., Rafiastanto, A., Rombang, W. M., and Trainor, C. R. (2002). Rediscovery of the critically endangered Sumatran Ground Cuckoo Carpococcyx viridis. Forktail, 18, 63–5.

Zuñiga, J. M. and Redondo, T. (1992b). Adoption of Great Spotted Cuckoo Clamator glandarius fledglings by Magpies Pica pica. Bird Study, 39, 200–2.

Zimmer, J. T. (1930). Birds of the Marshall Field Peruvian Expedition, 1922–1923. Field Museum of

Zwickl, D. J. and Hillis, D. M. (2002). Increased taxon sampling greatly reduces phylogenetic error. Systematic Biology, 51, 588–98.

Index Page numbers in bold indicate main species accounts and numbers in italic indicate illustrations and charts. A abbreviations xv acknowledgements v adaptive features, morphology 27 Africa, rains and breeding seasons 21 rains and migration 24 African Black Coucal 246–248, plate 5 breeding 248, description 246, displays and breeding behavior 248, field characters 247, food 248, habitat and general habits 247, measurements and weights 246, range and status 247, voice 247 African coucals plate 6 African Cuckoo 498–500 breeding 499, description 498, field characters 499, food 499, habitat and general habits 499, measurements and weights 498, range and status 499, voice 499 African Emerald Cuckoo 400–403, plate 11 breeding 402, description 400, displays and breeding behavior 402, field characters 401, food 402, geographic variation 401, glossy plumage 50, habitat and general habits 402, measurements and weights 401, range and status 402, voice 401 African Yellowbill, song 104 age, change to plumage 46 aggressive behavior, female 42 Alma de gato 331 American Striped Cuckoo 183–186, plate 1 breeding and life cycle 186, description 183, displays and breeding behavior 185, field characters 184, food 185, geographic variation 184, habitat and general habits 185, hooks on nestling’s bill 147, measurements and weights 184, range and status 184, voice 184 anatomical characters of cuckoos 66 Annam Ground-cuckoo 266 Anu Branco 169 arboreal and terrestrial cuckoos, differences 33, 34 relationships 58 arid habitat 11 Ash-colored Cuckoo 330–331, plate 9 breeding 331, description 330, field characters 330, food 331, habitat and general habits 331, measurements and weights 330, range and status 331, voice 331 Asian Drongo Cuckoo 460 Asian Emerald Cuckoo 385–388, plate 11 breeding 387, description 385, field characters 386, food 387, glossy plumage 50, habitat and general habits 387, history 386, measurements and weights 386, range and status 386, voice 386 Asian Koel 369 Asian Lesser Cuckoo 488–491, plate 16 breeding 490, description 488, field characters 488, food 490, habitat and general habits 490, measurements and weights 488, range and status 489, voice 489 Asian Violet Cuckoo, glossy plumage 50

asynchronus hatching 120 Australia, coucals plate 4 migration 24 rains and breeding season 23 Australian Bronze-cuckoo 405 Australian Koel 369 Australian Pheasant Coucal 257 breeding behavior 123 Azara’s Cuckoo 336 Azure-winged Magpie, changing response to parasitism 158 B Banded Bay Cuckoo 430–433, plate 13 breeding and life cycle 432, description 431, field characters 431, field characters 431, food 432, habitat and general habits 432, measurements and weights 431, range and status 432, subspecies 431, voice 431 Banded Ground-cuckoo 200–201, plate 2 breeding 201, description 200, field characters 201, habitat and general habits 201, measurements 200, range and status 201, voice 201 Barred Long-tailed Cuckoo 451–453, plate 14 breeding 453, description 451, field characters 452, food 453, habitat and general habits 453, measurements and weights 452, range and status 452, subspecies 451, voice 452 barriers to dispersal of species 106 Bay Coucal 221–222, plate 3 breeding 222, description 221, field characters 221, food 222, habitat and general habits 222, measurements 221, range and status 221, subspecies 221, voice 221 Bay-breasted Cuckoo 357 Bay-winged Cowbird, brood-parasitic behavior 164 beaters, using animals and other birds as 19 begging behavior, and calls 147 nestlings 93 begging calls, mimicking host call 149 behavior 15; see also individual species accounts Bernstein’s Coucal 255 Biak Coucal 213, plate 3 breeding 213, description 213, field characters 213, habitat and general habits 213, measurements 213, range and status 213, voice 213 bibliography 526 bill, measurements xx shape and color of 32 Black and White Cuckoo 320 Black-bellied Coucal 246 Black-bellied Cuckoo 335–336, plate 9 breeding 336, description 335, field characters 335, food 336, habitat and general habits 336, measurements and weights 335, range and status 335, voice 335

608 Index Black-bellied Malkoha 302–303, plate 8 breeding 303, description 302, displays and breeding behavior 303, field characters 303, food 303, habitat and general habits 303, measurements and weights 302, range and status 303, subspecies 302, voice 303 Black-billed Cuckoo 351–354, plate 9 begging behavior 93, breeding 354, description 351, displays and breeding behavior 353, field characters 351, food 353, habitat and general habits 353, measurements and weights 351, nesting 120, night migration 25, range and status 352, tent caterpillar diet 17, voice 352 Black-billed Koel 369 Black-capped Koel 367 Black Coucal 246 Black Crested Cuckoo 320 Black Cuckoo 481–485, plate 16 breeding 484, description 481, displays and breeding behavior 484, field characters 483, food 484, habitat and general habits 484, measurements and weights 483, range and status 483, subspecies 482, voice 483 Black-eared Cuckoo 407–409, plate 12 breeding 408, description 407, displays and breeding behavior 408, field characters 408, food 408, habitat and general habits 408, measurements and weights 408, range and status 408, voice 408 Black-faced Coucal 216–217, plate 3 breeding 217, description 216, field characters 217, food 217, geographic variation 217, habitat and general habits 217, measurements and weights 217, range and status 217, voice 217 Black-hooded Coucal 218–219, plate 3 breeding 219, description 218, field characters 218, food 219, habitat and general habits 219, measurements and weights 218, range and status 218, voice 218 Black Jungle Coucal 211 Black Scrub Coucal 255 Black-throated Coucal 224–225, plate 6 breeding 225, description 224, field characters 224, food 225, habitat and general habits 225, measurements and weights 224, range and status 225, subspecies 224, voice 224 Black Witch 174 Blue Coua 271–272, plate 7 breeding 272, description 271, displays and breeding behavior 272, field characters 272, following lemurs 19, food 272, habitat and general habits 272, measurements and weights 272, range and status 272, voice 272 Blue-faced Malkoha 301–302, plate 8 breeding 302, description 301, field characters 301, food 302, habitat and general habits 301, measurements and weights 301, range and status 301, voice 301 Blue-headed Coucal 228–230, plate 6 breeding 230, description 228, field characters 229, food 230, geographic variation 229, habitat and general habits 230, measurements and weights 229, range and status 230, subspecies 229, voice 229 Blue-headed Koel 369 Blue Malkoha 285

Blyth’s Cuckoo 505 body mass and wing length 33, 39 body weight 29 measurements xx Bornean Ground-cuckoo 264–266, plate 7 breeding 266, description 264, field characters 265, food 266, habitat and general habits 265, measurements and weights 264, range and status 265, voice 265 brain, size of 29, 32 in relation to body 28 brain-fever 117 breeding see individual species accounts breeding, cooperative 132 breeding displays 114 breeding seasons 21 and brood-parastic cuckoos 21 and calls 117 and rains 24 breeding success of host, effect of brood parasitism 150, 151 brooding 122 brood parasitism 3, 137, plate 18 behavior, observing 139 cost and benefits to hosts 154 effect on breeding success of host 150, 151 factor in reproductive success 159, 160 monophyly and skeletal characters 63 origins in New and Old World 164 single origin 62 brood-parasitic cuckoos, breeding season 21 incubation time 128 mating systems 118 size of brain 32 timing of migration 24 brood patch 128 Brown Coucal 238 Brush Cuckoo 439–447, plate 13 breeding and life cycle 446, description 439, displays and breeding behavior 446, field characters 444, food 446, habitat and general habits 445, history and subspecies 440, measurements and weights 442, range and status 445, voice 444 brush-cuckoos, songs 102 Buff-headed Coucal 208–209, plate 3 breeding 209, description 208, field characters 208, food 209, habitat and general habits 209, measurements and weights 208, range and status 209, subspecies 208, voice 209 C Cacomantis 421, plate 13 C. castaneiventris 426–427 C. flabelliformis 427–430 C. leucolophus 424–425 C. merulinus 433–436 C. pallidus 422–424 C. passerinus 437–439 C. sonneratii 430–433 C. variolosus 439–447 calls and breeding season 117 calls and songs 116; see also individual species accounts calls, begging 147

Index 609 Capsicum, dispersal of seeds by birds 20 captivity 14 Carpococcyx 262, plate 7 C. radiatus 264–266 C. renauldi 266–268 C. viridis 262–264 caterpillars as food 16 caterpillars, cleaning 16 Celebes Koel 369 Centropodidae 6, 70, 82, 208 phylogeny 85 Centropus 208, plate 4 C. anselli 222–223 C. ateralbus 210–211 C. bengalensis 250–254 C. bernsteini 255–257 C. celebensis 221–222 C. chalybeus 213 C. chlororhynchos 215–216 C. cupreicaudus 231–232 C. goliath 242–243 C. grillii 246–248 C. leucogaster 224–225 C. melanops 216–217 C. menbeki 211–213 C. milo 208–209 C. monachus 228–230 C. nigrorufus 236–237 C. phasianinus 257–262 C. rectunguis 219–220 C. senegalensis 226–228 C. sinensis 238–242 C. steerii 218–219 C. superciliosus 233–236 C. toulou 244–246 C. unirufus 214–215 lineage 85 C. violaceus 254–255 C.viridis 248–250 Cercococcyx 447, plate 14 C. mechowi 447–449 C. montanus 451–453 C. olivinus 449–451 cervical vertebrae 55 Ceuthmochares 285 C. aereus 287–290 C. australis 285–287 Ceylon Green-billed Coucal 215 Channel-billed Cuckoo 381–385, plate 10 breeding 384, description 381, displays and breeding behavior 384, field characters 382, food 384, habitat and general habits 384, measurements and weights 382, range and status 383, subspecies and history 381, voice 382 Chattering Yellowbill 287–290, plate 8 breeding 290, description 288, field characters 289, food 290, habitat and general habits 290, measurements and weights 288, range and status 289, subspecies and geographic variation 288, voice 289 Chestnut-bellied Cuckoo 356–357, plate 9 breeding 357, description 356, field characters 356, food 357, habitat and general habits 356,

measurements and weights 356, range and status 356, voice 356 Chestnut-bellied Malkoha 298–299, plate 8 breeding 299, description 298, field character 298, food 299, habitat and general habits 298, history 298, measurements and weights 298, range and status 298,voice 298 Chestnut-breasted Cuckoo 426–427, plate 13 breeding 427, description 426, displays and breeding behavior 427, field characters 426, food 427, habitat and general habits 427, measurements and weights 426, range and status 427, subspecies 426, voice 427 Chestnut-breasted Malkoha 294–297, plate 8 breeding 297, description 295, field characters 296, food 297, habitat and general habits 297, measurements and weights 296, range and status 296, subspecies 295, voice 296 Chestnut-winged Crested Cuckoo 311 Chestnut-winged Cuckoo 311–313, plate 15 breeding 313, description 311, field characters 311, food 312, habitat and general habits 312, measurements and weights 311, range and status 312, voice 311 chromosomes 69 Chrysococcyx 385, plates 11, 12 C. basalis 405–407 C. caprius 391–395 C. cupreus 400–403 C. flavigularis 399–400 C. klaas 395–399 C. lucidus 410–413 C. maculatus 385–388 C. megarhynchus 403–404 C. meyeri 413–414 C. minutillus viii, 415–421 C. osculans 407–409 C. ruficollis 409–410 C. xanthorhynchus 388–391 Clamator 311, plate 15 C. coromandus 311–313 C. glandarius 313–317 C. jacobinus 320–325 C. levaillantii 318–320 clutches 121 Coccycua 91, 325, plate 9 C. cinerea 330–331 C. minuta 325–328 C. pumila 328–329 Coccyzidae 70 Coccyzus 336, plate 9 C. americanus 339–344 C. erythropthalmus 351–354 C. euleri 345–347 C. ferrugineus 350–351 C. lansbergi 354–356 C. longirostris 362–363 C. melacoryphus 336–339 C. merlini 361–362 C. minor 347–349 C. pluvialis 356–357 C. rufigularis 357–358 C. vetula 358–359 C. vieilloti 360–361

610 Index Cocos Cuckoo 350–351, plate 9 breeding 351, description 350, field characters 350, food 351, habitat and general habits 350, measurements 350, range and status 350, voice 350 coevolution of host and brood-parasite species 107, 155, 156 color reflectance 50 color and pattern of eggs, matching hosts’ eggs 143, 153, 155 color of plumage xx Common Coucal (India), 238 Common Coucal (New Guinea), 257 Common Cuckoo 510–517, plate 16 begging behavior 148, breeding 516, changing choice of host 156, courtship 115, description 510, displays and breeding behavior 515, effect on host 151, field characters 512, food 515, habitat and general habits 515, history and subspecies 511, measurements and weights 512, migration 24, mimicking egg color and pattern 142, morph in egg color 156, range and status 513, voice 513 Common Hawk-cuckoo 471–473, plate 15 breeding 473, description 471, field characters 472, food 473, habitat and general habits 473, measurements and weights 472, range and status 472, subspecies 472, voice 472 Common Koel 369–378, plate 10 breeding and life cycle 376, description 369, displays and breeding behavior 376, field characters 373, first brood parasite 137, food 376, habitat and general habits 375, measurements and weights 372, plumage phases 49, range and status 374, subspecies 369, systematic notes 370, voice 373 Common Roadrunner, vocal repertoire 117, comparative data, cuckoo species 35 comparative features, morphology 27 Congo Yellowbill 287 conservation xxi conservation status 11; see also individual species accounts conspecific mating 145 Cooee 369 cooperative breeding 132, 162 Coppery-tailed Coucal 231–232, plate 6 breeding 232, description 231, field characters 231, food 232, habitat and general habits 232, measurements and weights 231, range and status 231, subspecies 231, voice 231 copulation 121 Coquerel’s Coua 276–277, plate 7 breeding 277, description 276, displays and breeding behavior 277, field characters 276, food 277, habitat and general habits 277, measurements and weights 276, range and status 276, voice 276 Coral-billed Ground-cuckoo 266–268, plate 7 breeding 268, description 266, field characters 267, food 267, habitat and general habits 267, history 266, measurements and weights 266, range and status 267, voice 267 Coua 268 C. caerulea 271–272 C. coquereli 276–277

C. cristata 268–270 C. cursor 277–278 C. delandei 280–281 C. gigas 279–280 C. reynaudii 274–275 C. ruficeps 273–274 C. serriana 281–282 C. verreauxi 270–271 couas 7 coucals 6 diet 17 long hallux claw 41 Couinae 7, 82, 262 phylogeny 87 courtship feeding 115, 116 Crested Coua 268–270, plate 7 breeding 270, description 268, displays and breeding behavior 270, field characters 269, food 270, habitat and general habits 270, measurements and weights 269, range and status 269, subspecies 269, voice 269 Crotophaga 172, plate 1 C. ani 174–178 C. major 172–174 C. sulcirostris 178–183 Crotophagidae 6, 70, 82, 169 phylogeny 84 crotophagines 6 Crow-Pheasant 238 Cuban Lizard-cuckoo 361–362, plate 9 breeding 362, description 361, field characters 361, food 362, habitat and general habits 362, measurements and weights 361, range and status 362, subspecies 361, voice 361 cuckoldry 3, 137 Cuckoo 510 cuckoos as food 13 as medicine 13 cuckoo species, comparative data 35 Cuculinae 7, 82, 283 cuculines 7 Cuculini, phylogeny 90 Cuculus 481, plate 16 C. canorus 510–517 C. clamosus 481–485 C. crassirostris 491–492 C. gularis 498–500 C. lepidus 508–510 C. micropterus 492–496 C. optatus 500–505 C. poliocephalus 488–491 C. rochii 496–498 C. saturatus 505–508 C. solitarius 485–488 D Dark-backed Bronze-cuckoo 415 Dark-billed Cuckoo 336–339, plate 9 breeding 338, description 337, displays and breeding behavior 338, field characters 337, food 338, habitat and general habits 338, measurements and weights 337, range and status 337, voice 337

Index 611 Dark Hawk-cuckoo 466–468, plate 15 breeding 468, description 466, field characters 467, food 468, habitat and general habits 467, measurements and weights 467, range and status 467, voice 467 Dasylophus 308 D. cumingi 309–311 D. superciliosus 308–309 daylight and breeding 21 defensive behavior of hosts 152 Delalande’s Coua, extinction 13, plate 7 description 5; see also individual species accounts Didric Cuckoo 391 Diederik Cuckoo 391–395, plate 11 breeding 394, courtship feeding 116, description 391, displays and breeding behavior 393, field characters 392, food 393, geographic variation 391, habitat and general habits 393, mating 118, measurements and weights 392, molt 392, range and status 392, voice 392 dimorphism, male and female sizes 41, 42, 43 display behavior 114 distribution 9, 105; see also individual species accounts DNA sequencing, specimens used in author study 73–80 DNA studies 69, 70 Dove Cuckoo 500 Dromococcyx 187, plate 1 D. pavonius 189–191 D. phasianellus 187–189 drongo cuckoos 12, 460 species and songs 100 Dusky Long-tailed Cuckoo 447–449, plate 14 breeding 449, description 447, field characters 448, 449, habitat and general habits 449, measurements and weights 448, range and status 448, voice 448 Dwarf Cuckoo 328–329, plate 9 breeding 329, description 328, displays and breeding behavior 329, field characters 328, food 329, habitat and general habits 329, measurements and weights 328, range and status 328, voice 328 Dwarf Koel 367–368, plate 10 breeding 368, description 367, field characters 368, food 368, habitat and general habits 368, measurements and weights 367, range and status 368, subspecies 367, voice 368 E East Africa, rains and breeding season 21 ecological influences on host-cuckoo relationship 154 eggs 121, 123, plate 17 color and pattern 142, 143, 153, 155 food 16 problems in identification 138 rounded shape 126 shape 125 size and eggshell weight 127 size and female body size 124 size and nestling period 129 sizes and shapes 123 egg-laying in host nest 139 eggshell 126 thick outer layer 127 weight and size of eggs 127

embryo, development 128 Emerald Cuckoo 400 Eudynamys 368, plate 10 E. solopacea 369–378 Eurasian Cuckoo 510 European Cuckoo 510 European and North American host birds, differences 153 eviction behavior of nestlings 145, 146 evolution 111 brood parasitism, breeding behavior 161 evolutionary history 4 evolutionary relationships among cuckoos, author study 70, 71, 72 eyelashes 43 F face and head, feathers 44 facultative brood parasitism 162 Fan-tailed Brush Cuckoo 427 Fan-tailed Cuckoo 427–430, plate 13 breeding 430, description 428, displays and breeding behavior 430, field characters 429, food 430, habitat and general habits 430, history 428 measurements and weights 428, range and status 429, subspecies 428, voice 429 feathers 5 face and head 44 structural features 50 feather tracts, morphological variations 65 feeding young 122 field characters see individual species accounts figs, calcium source 19 fledging 122 fledglings, size and parental behavior 130 flight 5 flight feathers, molting 52 “flush and rush” 17 food and feeding 16; see also individual species accounts forest cuckoos, at risk from encroachment 11, 12 Fork-tailed Drongo-cuckoo 454–457, plate 14 breeding 457, description 454, field characters 455, food 457, habitat and general habits 457, measurements and weights 455, range and status 456, species and subspecies 454, voice 455 Foro panarium, earliest fossil 109 fossil evidence 109 foster parents 137 mimicking songs 118 fruit in diet 19 G Gabon Coucal 222–223, plate 6 breeding 223, description 222, field characters 223, food 223, habitat and general habits 223, measurements and weights 223, range and status 223, voice 223 genes 143 genes and species 96 genetic influences on host-cuckoo relationship 154 gens theory 141

612 Index Geococcyx 193, plate 2 G. californianus 193–198 G. velox 198–200 geographic speciation 105 geology and species distribution 106, 107 Giant Coua 279–280, plate 7 breeding 280, description 279, field characters 279, food 280, habitat and general habits 279, measurements and weights 279, range and status 279, voice 279 Giant Coucal 242 Giant Forest Coucal 254 Golden Bronze-cuckoo 410 Goliath Coucal 242–243, plate 5 breeding 243, description 242, field characters 243, food 243, habitat and general habits 243, measurements 243, range and status 243, voice 243 Gould’s Bronze-cuckoo 415 Great Roadrunner, pelvic muscles 35 Great Spotted Cuckoo 313–317, plate 15 breeding 316, description 313, displays and breeding behavior 315, field characters 314, food 315, habitat and general habits 315, host species 108, measurements and weights 313, range and status 314, subspecies and history 313, voice 314 Greater Ani 172–174, plate 1 breeding 174, description 172, field characters 173, following ants and squirrel monkey troops 19, food 174, habitat and general habits 173, measurements and weights 172, range and status 173, voice 173 Greater Black Coucal 211–213, plate 3 breeding 212, description 211, field characters 212, food 212, habitat and general habits 212, measurements and weights 212, range and status 212, subspecies 211, voice 212 Greater Coucal 211, 238–242, plate 4 breeding 241, description 238, displays and breeding behavior 241, field characters 239, food 241, habitat and general habits 240, measurements and weights 239, range and status 240, subspecies 238, voice 240 Greater Roadrunner 193–198, plate 2 breeding and life cycle 197, breeding display 114, description 193, displays and breeding behavior 197, facial feathers 45, field characters 194, food 196, food for young 122, habitat and general habits 195, history 194, human impact14, killing a rattlesnake 18, measurements and weights 194, range and status 195, voice 194 Green-billed Coucal 215–216, plate 4 breeding 216, breeding behavior 216, description 215, field characters 215, food 216, habitat and general habits 216, measurements 215, range and status 215, voice 215 Green-billed Ground-cuckoo 262, 264 Green-billed Malkoha 304–306, plate 8 breeding 306, description 304, displays and breeding behavior 305, field characters 305, food 305, habitat and general habits 305, measurements and weights 304, range and status 305, subspecies and taxonomic comments 304, voice 305 Green Coucal 285, 287 Green Malkoha 285, 287

Grey-bellied Cuckoo 437–439, plate 13 breeding and life cycle 438, description 437, displays and breeding behavior 438, field characters 438, food 438, habitat and general habits 438, measurements 437, range and status 438, systematics 437, voice 438 Grey-breasted Brush Cuckoo 439 Grey Bushchat 156 Grey-capped Cuckoo 354–356, plate 9 breeding 356, description 354, field characters 355, food 355, habitat and general habits 355, measurements and weights 355, range and status 355, voice 355 Grey Cuckoo 510 Grey Gerygone, cost of Shining Bronze-cuckoo parasitism 151 Grey-headed Cuckoo 439 Groove-billed Ani 178–183, plate 1 begging behavior 93, breeding 134, 181, description 178, diet 17, displays and breeding behavior 180, field characters 179, food 180, habitat and general habits 180, measurements and weights 179, range and status 179, voice 179 ground-cuckoos 6, 7 feeding on ants 18 Guira 169, plate 1 G. guira 169–174 Guira Cuckoo 169–174, plate 1 breeding and life cycle 133, 171, cooperative breeding 162, description 169, displays and breeding behavior 171, facial feathers 45, field characters 169, food 171, habitat and general habits 170, measurements and weights 169, other names 169, range and status 170, social behavior 135, sunning behavior 11, voice 170 gypsy moths as food 17 H habitat see individual species accounts habitats 10; see also individual species account hackles 45 hallux claw, long 41 sexual size dimorphism 44 hatching 121 hawk-cuckoos, species and songs 98, 99 head and bill, shape of 32 helper birds 132, 133 Hierococcyx 464, plate 15 H. bocki 466–468 H. fugax 477–479 H. hyperythrus 473–475 H. nisicolor 479–481 H. pectoralis 475–477 H. sparverioides 468–471 H. vagans 465–466 H. varius 471–473 Himalayan Cuckoo 505–508, plate 16 breeding 508, description 505, field characters 506, food 508, habitat and general habits 507, measurements and weights 506, range and status 506, voice 506 Himalayan Hawk-cuckoo 479 Hispaniolan Lizard-cuckoo 362–363, plate 9 breeding 363, description 362, field characters 363, food 363, habitat and general habits 363,

Index 613 measurements and weights 363, range and status 363, subspecies 363, voice 363 Hoatzin, relationship to cuckoos 112 Hodgson’s Hawk-cuckoo 479 hooks on nestling’s bill 147 Horsfield’s Bronze-cuckoo 405–407, plate 12 begging behavior 93, begging call 148, breeding 407, description 405, displays and breeding behavior 407, eviction behavior of nestling 144, field characters 405, food 406, habitat and general habits 406, measurements and weights 405, nestlings 146, range and status 406, voice 405 Horsfield’s Cuckoo 500 Horsfield’s Hawk-cuckoo 473 host egg, mimicry 137, 142 host nest, finding 139 host, genetic influences 154 mimicking call 148, 149 mimicking egg color 142 host-choice behavior, changes in 156 host-parasite cospeciation 107 host species 13 effect of brood parasitism 150, 151 suitability criteria 140, 141 switching 108 host-specific begging call 148, 149 host-tolerant species 147 Hume’s Crow-Pheasant 238 humerus, cuckoos and other birds 57 Hunter 356 I, J imprinting 140 incubation 127 period 129 period and egg size 128, India, altitude and breeding season 22 Indian Cuckoo 492–496, plate 16 breeding 495, description 492, displays and breeding behavior 495, field characters 494, food 495, habitat and general habits 495, measurements and weights 493, range and status 494, subspecies and history 493, voice 494 Indian Plaintive Cuckoo 437 Indonesian Cuckoo 439 insect population 13 insects as food 16 intraspecific brood parasitism 162 Jacobin Cuckoo 320–325, plate 15 breeding 324, call 116, color phases 49, courtship feeding 115, description 320, displays and breeding behavior 323, field characters 321, food 323, habitat and general habits 323, history 321, measurements and weights 321, range and status 321, voice 321 Jamaican Lizard-cuckoo 358–359, plate 9 breeding 359, description 358, displays and breeding behavior 359, field characters 359, food 359, habitat and general habits 359, measurements and weights 358, range and status 359, voice 359 Javan Coucal 236–237, plate 4 breeding 237, description 236, field characters 236, food 237, habitat and general habits 237, measurements 236, voice 237

Javan Hawk-cuckoo 477–479, plate 15 breeding 479, description 477, field characters 477, food 479, habitat and general habits 478, measurements and weights 477, range and status 478, systematic notes 477, voice 478 Jungle Babblers, host to Jacobin Cuckoo and Common Hawk-Cuckoo 140 juvenile plumage 46 K, L Kai Coucal 257 killing and mobbing of cuckoos 152 killing host young 150 Klaas’s Cuckoo 395–399, plate 11 breeding and life cycle 398, description 395, displays and breeding behavior 398, field characters 397, fledgling 16, food 398, habitat and general habits 397, measurements and weights 396, molt 396, range and status 397, voice 397 Koels, figs in diet 19 Large Hawk-cuckoo 468–471, plate 15 breeding 470, description 468, field characters 469, food 470, habitat and general habits 470, measurements and weights 468, range and status 469, voice 469 Lark-heeled Cuckoo 238 leaves in nestbuilding 121 leg length 29 leg length, differences in arboreal and terrestrial cuckoos 33, 34 leg muscles, morphological variations 65 lemurs, birds following 19 Lesser Black Coucal 255–257, plate 4 breeding 257, description 255, field characters 256, food 256, habitat and general habits 256, measurements and weights 256, range and status 256, subspecies 256, voice 256 Lesser Coucal 250–254, plate 5 breeding 254, description 250, displays and breeding behavior 253, field characters 252, food 253, habitat and general habits 253, measurements and weights 252, range and status 253, subspecies 251, voice 252 Lesser Green-billed Malkoha 301 Lesser Ground-cuckoo 191–193, plate 1 breeding and life cycle 193, description 191, field characters 192, food 193, habitat and general habits 193, measurements and weights 192, other names 191,range and status 192, subspecies and geographic variation 192, voice 192 Lesser Roadrunner 198–200, plate 2 breeding and life cycle 200, description 198, field characters 199, food 200, habitat and general habits 199, history and variation 199, measurements and weights 198, range and status 199, voice 199 Levaillant’s Cuckoo 318–320, plate 15 breeding 319, color phases 49, description 318, displays and breeding behavior 319, field characters 318, food 319, habitat and general habits 319, measurements and weights 318, range and status 318, voice 318 light waves, reflected from feathers 50 lineage, basal 83 lineages, origins of 83, 84

614 Index Linnaeus 4 Little Bronze-cuckoo 415–421, plate 12 breeding 421, description 415, displays and breeding behavior 421, field characters 419, food 421, habitat and general habits 420, measurements and weights 419, range and status 420, relationship with flyeaters 107, song 103, subspecies 415, voice 420 Little Cuckoo 325–328, plate 9 breeding 327, description 325, field characters 326, food 327, habitat and general habits 327, measurements and weights 326, range and status 327, subspecies 326, voice 327 Little Malkoha 283 lizards as food 16 logging 11 Long-billed Cuckoo 403–404, plate 12 breeding 404, description 403, field characters 404, food 404, habitat and general habits 404, measurements and weights 404, range and status 404, voice 404 Long-tailed Cuckoo 378–380, plate 10 breeding 380, description 378, displays and breeding behavior 380, field characters 379, food 380, habitat and general habits 380, measurements and weights 379, range and status 379, voice 379 Long-tailed Koel 378 M Madagascar Coucal 244–246, plate 5 breeding 245, description 244, displays and breeding behavior 245, field characters 244, food 245, habitat and general habits 245, measurements and weights 244, range and status 245, subspecies 244, voice 245 Madagascar Cuckoo 496 Madagascar Lesser Cuckoo 496–498, plate 16 breeding 498, description 496, field characters 497, food 498, habitat and general habits 498, measurements and weights 497, range and status 497, voice 497 Madenfresser 17 Magpie, changing response to parasitism 157 Malay Bronze-cuckoo 415 malkohas 7 malkohas, Old World, phylogeny 89 Mangrove Cuckoo 347–349, plate 9 breeding 349, description 347, displays and breeding behavior 349, field characters 348, food 349, habitat and general status 348, measurements and weights 347, range and status 348, subspecies 347, voice 348 maps see individual species accounts mating systems 118 Maynard’s Cuckoo 347 Meadow Pipit, effect of Common Cuckoo parasitism 152 measurements xx; see also individual species accounts Menbek’s Coucal 211 Meyer’s Bronze-cuckoo 413 Microdynamis 367, plate 10 M. parva 367–368 migrants, vagrants 25 migration 23 difficulties in recording 25 and rains 24

mimetic eggs 143, 154, 155 mimicking hosts’ egg color 142, 153, 155 songs of foster parents 118 mimicry, genetics of 142 host egg 137 mitrochondrial DNA studies 70, 71 mixed-species flocks, feeding in 19 mobbing and killing of cuckoos 152 molting 51 molt sequences 52 Moluccan Brush Cuckoo 439 Moluccan Drongo-cuckoo 463–464, plate 14 breeding 464, description 463, field characters 463, habitat and general habits 464, measurements and weights 463, range and status 464, systematic comments 463, voice 464 monogamy 119 monophyly of cuckoos 81 monophyly and skeletal characters 60 Morococcyx 191, plate 1 M. erythropygus 191–193 morphological variations 65 morphology 27; see also individual species accounts Mountain Bronze-cuckoo 409, 413 Mountain Cuckoo 505 mouth patterns, young cuckoos plate 20 Mustached Hawk-cuckoo 465–466, plate 15 breeding 466, description 465, field characters 465, food 466, habitat and general habits 466, measurements and weights 465, range and status 466, voice 465 N Narrow-billed Bronze-cuckoo 405 natal down 45 Neomorphidae 6, 70, 82, 183 Neomorphinae 6 Neomorphus 200, plate 2 N. geoffroyi 204–207 N. pucheranii 203–204 N. radiolosus 200–201 N. rufipennis 202–203 nest predation 159 factor in brood-parasitism strategies 160 nesting hosts, mobbing cuckoos 152 nesting pairs 121 nestlings, development 129 foster parents’ attitudes to 153 period and egg size 129 nests and nestbuilding 120 nests, takeover 163 New Guinea Pheasant Coucal 257 New World brood parasites, origins 164, 165 New World cuckoos, phylogeny 84 New World tropics, habitat 10 night migration 25 nomenclature xix North American and European host birds, differences 153 North American species, migration 23 Northern Hawk-cuckoo 473

Index 615 Northern Muted Cuckoo 500 nostrils, shape of 32 O obligate brood parasites 159 Old Man Bird 356 Old Woman Bird 358 Old World brood parasites, origins 164, 165 Old World species, plates 7, 8, 10 migration 24 tropics, habitat 10 Olive Long-tailed Cuckoo 449–451, plate 14 breeding 451, description 449, field characters 450, food 451, habitat and general habits 450, measurements and weights 450, range and status 450, voice 450 onomatopoeic calls 118 Opisthocomidae 70 Oriental Cuckoo 500–505, plate 16 breeding 505, description 500, field characters 502, food 505, habitat and general habits 504, history 501, measurements and weights 502, range and status 503, voice 503 P Pachycoccyx 364, plate 10 P. audeberti 364–366 Pacific Long-tailed Cuckoo 378 Pallid Cuckoo 422–424, plate 13 breeding 424, description 422, displays and breeding behavior 424, field characters 423, food 424, habitat and general habits 424, history 423, measurements and weights 423, plumage variation 49, range and status 423, voice 423 parasitic breeding behavior 3 parental behavior 15 parental care 159 Pavonine Cuckoo 189–191, plate 1 breeding 191, description 189, field characters 190, food 191, habitat and general habits 190, measurements and weights 190, range and status 190, voice 190 Peacock Cuckoo 189 Pearly-breasted Cuckoo 345–347, plate 9 breeding 347, description 345, field characters 345, food 347, habitat and general habits 346, measurements and weights 345, range and status 346, voice 345 peccaries and wild pigs, birds following 19 “peer and pounce” 16 Phaenicophaeini 88, 89 Phaenicophaeus 294 P. curvirostris 294–297 P. diardi 302–303 P. pyrrhocephalus 299–301 P. sumatranus 298–299 P. tristis 304–306 P. viridirostris 301–302 Pheasant Coucal 257–262, plate 4 breeding 261, description 257, displays and breeding behavior 261, field characters 260, food 261, habitat and general habits 261, history and subspecies 258, measurements and weights 259, plumage 49, range and status 260, voice 260

Pheasant Cuckoo 187–189, plate 1 breeding 189, description 187, displays and breeding behavior 189, field characters 188, food 189, habitat and general habits 188, measurements and weights 187, other names 187, range and status 188, subspecies 187, voice 188 Philippine Coucal 248–250, plate 5 breeding 250, description 248, field characters 249, food 250, habitat and general habits 249, measurements and weights 249, range and status 249, subspecies 249, voice 249 Philippine Drongo-cuckoo 458–459, plate 14 breeding 459, description 458, field characters 459, food 459, habitat and general habits 459, measurements and weights 458, range and status 459, subspecies and systematic comments 458, voice 459 Philippine Hawk-cuckoo 475–477, plate 15 breeding 477, description 475, field characters 476, food 476, habitat and general habits 476, measurements and weights 476, range and status 476, voice 476 phylogenetic analysis 69, 70 phylogeny xxi and behavior models of brood parasitism 161 and lineages of brood-parasitic birds 163 Centropodinae 85 Couinae 87 Crotophaginae 84 Cuculini 90 genus-level 82 Neomorphinae 84 New World cuckoos 84 Old World malkohas 89 Piaya 331, plate 9 P. cayana 331–334 P. melanogaster 335–336 Pied Bronze-cuckoo 415 Pied Coucal 210–211, plate 3 breeding 211, description 210, field characters 210, food 211, habitat and general habits 211, measurements and weights 210, range and status 210, voice 210 Pied Crested Cuckoo 320 Pied Cuckoo 320 pig birds 19 Piririgua 169 Plaintive Cuckoo 433–436, plate 13 breeding and life cycle 436, description 433, displays and breeding behavior 436, field characters 435, food 436, habitat and general habits 436, measurements and weights 435, range and status 435, subspecies 433, systematics 433, voice 435 plaintive-cuckoos, songs 101 plumage 43 change with age 46 color xx color morphs 48, 49 glossy 49 juvenile 46 sexual dimorphism 48 subadult 47

616 Index polyandry 119 prance display 114 predation 159, 160 predators 20 avoidance tactics 20 Puerto Rican Lizard-cuckoo 360–361, plate 9 breeding 361, description 360, field characters 360, food 361, habitat and general habits 360, measurements and weights 360, range and status 360, voice 360 R Radiated Fruit-cuckoo 264 Raffles’s Malkoha 283–285, plate 8 breeding 285, description 283, field characters 284, food 285, habitat and general habits 285, measurements and weights 284, range and status 284, systematic comments 283, voice 284 Raffles, Stamford 4 rain birds 21 Rain Bird 369 rains and breeding season 21 and migration 24 range see individual species accounts recognizing own species 145 Red-billed Ground-cuckoo 203–204, plate 2 breeding 204, description 203, field characters 204, food 204, habitat and general habits 204, measurements and weights 204, range and status 204, subspecies 203, voice 204 Red-billed Malkoha 293–294, plate 8 breeding 294, description 293, field characters 293, food 294, habitat and general habits 294, measurements and weights 293, range and status 293, subspecies 293, voice 293 Red-breasted Coua 281–282, plate 7 breeding 282, description 281, field characters 282, food 282, habitat and general habits 282, measurements and weights 282, range and status 282, voice 282 Red-capped Coua 273–274, plate 7 breeding 274, description 273, field characters 273, food 274, habitat and general habits 274, measurements and weights 273, range and status 274, subspecies 273, voice 273 Red-chested Cuckoo 485–488, plate 16 breeding and life cycle 487, description 485, displays and breeding behavior 487, field characters 486, food 486, habitat and general habits 486, history and subspecies 485, measurements and weights 486, range and status 486, voice 486 Reddish-throated Bronze-cuckoo 409 Red-faced Malkoha 299–301, plate 8 breeding 300, breeding behavior 300, description 299, field characters 300, food 300, habitat and general habits 300, measurements 299, range and status 300, voice 300 Red-fronted Coua 274–275, plate 7 breeding 275, description 274, displays and breeding behavior 275, field characters 275, food 275, habitat and general habits 275, measurements and weights 275, range and status 275, voice 275 Red-winged Crested Cuckoo 311

Reed Warbler, cost of Common Cuckoo parasitism 151 rejecting cuckoo eggs 153, 154, 155 relationships to other birds 111 relatives of cuckoos 111 Renauld’s Ground-cuckoo 266 reproduction, cost of 161 reproductive success and brood parasitism 159, 160 Reynaud’s Coua 274 Rhamphococcyx 306 R. calyorhynchus 306–308 Rhinortha 283 R. chlorphaea 283–285 rivers, barriers to species dispersal 106 Roadrunners, diet 17 egg-laying 121 nests 122 parental behavior 122, Rough-crested Cuckoo 308–309, plate 8 breeding 309, description 308, field characters 309, food 309, habitat and general habits 309, measurements and weights 308, range and status 309, subspecies 308, voice 309 rounded shape, eggs 126 Rufous-bellied Coucal 226 Rufous-bellied Malkoha 298 Rufous-breasted Cuckoo 357–358, plates 4, 9 breeding 358, description 357, field characters 357, food 358, habitat and general habits 358, measurements and weights 357, range and status 357, voice 357 Rufous Bronze-cuckoo 415 Rufous Coucal 214–215, plate 3 breeding 215, description 214, field characters 214, food 214, habitat and general habits 214, measurements and weights 214, range and status 214, voice 214 Rufous Hawk-cuckoo 473–475, plate 15 breeding 475, description 473, field characters 474, food 475, habitat and general habits 475, measurements and weights 474, range and status 474, voice 474 Rufous-rumped Cuckoo 191 Rufous-tailed Bronze-cuckoo 405 Rufous-throated Bronze-cuckoo 409–410, plate 12 breeding 410, description 409, field characters 409, food 410, habitat and general habits 410, measurements and weights 409, range and status 409, voice 409 Rufous-vented Ground-cuckoo 204–207, plate 2 breeding 207, description 204, field characters 206, food 207, habitat and general habits 207, measurements and weights 206, range and status 206, subspecies 205, voice 206 Rufous-winged Ground-cuckoo 202–203, plate 2 breeding 203, description 202, field characters 202, food 203, habitat and general habits 203, measurements and weights 202, range and status 202, voice 202 Running Coua 277–278, plate 7 breeding 278, description 277, field characters 278, food 278, habitat and general habits 278, measurements and weights 278, range and status 278, voice 278 Rusty-breasted Cuckoo 439

Index 617 S Salvadori,Tommaso 4 Savanna Blackbird 174 Scale-feathered Malkoha 309–311, plate 8 breeding 310, description 309, field characters 310, food 310, habitat and general habits 310, measurements and weights 310, range and status 310, voice 310 Senegal Coucal 226–228, plate 6 breeding and life cycle 228, description 226, displays and breeding behavior 228, feeding on locusts 17, field characters 227, food 228, geographic variation 226, habitat and general habits 227, measurements and weights 226, range and status 227, subspecies 226, voice 227 Senegal Yellowbill 287 separation of major groups, timing 111 sexual differences 42 sexual size dimorphism 41, 42, 43 Shining Bronze-cuckoo 410–413, plate 12 begging call 148, breeding 413, description 410, displays and breeding behavior 413, field characters 411, food 412, habitat and general habits 412, measurements and weights 411, range and status 412, subspecies 411, voice 411 Shining Cuckoo 410 Short-toed Coucal 219–220, plate 3 breeding 220, description 219, field characters 219, food 220, habitat and general habits 220, measurements and weights 219, range and status 219, voice 219 single origin of brood parasitism 62 Sirkeer Cuckoo 291 Sirkeer Malkoha 291–292, plate 8 breeding 292, description 291, displays and breeding behavior 292, field characters 291, food 292, habitat and general habits 292, measurements and weights 291, range and status 292, subspecies 291, voice 292 size of species 6 skeletal characters and brood parasitism monophyly 63 skeletal characters and cuckoo monophyly 60 skeletal features 53 unique to cuckoos 57, 59 Small Green-billed Malkoha 301 Smooth-billed Ani 174–178, plate 1 breeding and life style 134, 177, description 174, displays and breeding behavior 177, feeding on butterflies 17, field characters 175, food 177, habitat and general habits 176, measurements and weights 175, other names 174, range and status 175, voice 175 Snail-eating Coua 280–281 breeding 281, description 280, extinction 13, field characters 280, food 281, habitat and general habits 281, measurements 280, range and status 281, voice 280 snails as food 17 social behavior 15 solitary nature of cuckoos 15 song and species 95, 96 songs and calls 116; see also individual species accounts similarities across species 97 South American species, migration 23

South-central Africa, breeding seasons 23 Southern Masked Weaver, cost of Diederik Cuckoo parasitism 152 speciation 105 species, at risk 12 descriptions xxi; see individual species accounts definition 95 Splendid Fairy-wren, cost of Horsfield’s Bronze-cuckoo parasitism 151 host to Shining Bronze-cuckoo 140 Square-tailed Cuckoo 439 Square-tailed Drongo-cuckoo 460–463, plate 14 breeding 462, description 460, displays and breeding behavior 462, field characters 461, food 462, habitat and general habits 462, measurements and weights 461, range and status 461, subspecies 460, voice 461 Squirrel Cuckoo 331–334, plate 9 breeding 334, call 116, description 332, displays and breeding behavior 334, field characters 333, food 334, habitat and general habits 334, history and subspecies 332, measurements and weights 333, range and status 333, voice 333 Sri Lanka Green-billed Coucal 215 status see individual species accounts Steere’s Coucal 218 sternum 56 Sualwesi Koel 369 subadult plumage 47 subspecies see individual species accounts recognition of xxi Sulawesi Cuckoo 491–492, plate 15 breeding 492, description 491, field characters 492, food 492, habitat and general status 492, measurements 491, range and status 492, voice 492 Sumatran Ground-cuckoo 262–264, plate 7 breeding 264, description 262, field characters 263, food 264, habitat and general habits 264, measurements 263, range and status 263, voice 263, vulnerable status 13 Sunda Coucal 236 Sunda Cuckoo 508 Sunda Ground-cuckoo 262, 264 Sunda Lesser Cuckoo 508–510 breeding 509, description 508, field characters 509, food 509, habitat and general habits 509, measurements 509, range and status 509, subspecies 509, voice 509 sunning behavior 11 Surniculus 453, plate 14 S. dicuroides 454–457 S. lugubris 460–463 S. musschenbroeki 463–464 S. velutinus 458–459 Swamp Pheasant 257 switching host species 108 syrinx, morphological variations 65 Scythrops 380, plate 10 S. novaehollandiae 381–385 T Taccocua 291 T. leschenaultii 291–292

618 Index tail 5 as rudder 45 tail feathers, measurements xx tail-wag display 114 Tapera 183, plate 1 T. naevia 183–186 tarsometatarsus, cuckoos and other birds 54, 56 taxonomy 68 tent caterpillars as food 16 terrestrial and arboreal cuckoos, differences 33, 34 relationships 58 Thick-billed Cuckoo 364–366, plate 10 breeding 366, description 364, displays and breeding behavior 366, field characters 365, food 366, habitat and general habits 365, measurements and weights 364, range and status 365, relationship with helmetshrikes 108, subspecies 364, voice 365 threats to species 11, 12 Tick Bird 174 tick-eaters 17 turacos, similarities to cuckoos 112 U,V Urodynamis 378, plate 10 U. taitensis 378–380 Verreaux’s Coua 270–271, plate 7 breeding 271, description 270, field characters 271, food 271, habitat and general habits 271, measurements 271, range and status 271, voice 271 vertebrates as food 17 Vidua finches, brood-parasitic behavior 153 Village Weaver, eggs plate 19 changing response to parasitism 157 Violaceous Coucal 254–255, plate 5 breeding 255, description 254, field characters 255, food 255, habitat and general habits 255, measurements and weights 254, range and status 255, voice 255 Violet Cuckoo 388–391, plate 11 breeding 390, description 388, displays and breeding behavior 390, field characters 389, food 390, habitat and general habits 390, measurements and weights 389, range and status 389, subspecies 388, voice 389 voice see individual species accounts W, X,Y, Z weights see individual species accounts West Africa, rains and breeding season 22 Western Thornbill, host to Shining Bronze-cuckoo 140 Whistling Hawk-cuckoo 479–481, plate 15 breeding 481, description 479, field characters 480, food 481, habitat and general habits 481, measurements and weights 480, range and status 480, voice 480

Whistling Yellowbill 285–287, plate 8 breeding 287, description 285, displays and breeding behavior 287, field characters 286, food 287, habitat and general habits 287, measurements and weights 286, range and status 286, taxonomic comments and geographic variation 286, voice 286 White Ani 169 White-bellied Didric Cuckoo 391 White-browed Coucal 233–236, plate 6 breeding and life cycle 235, description 233, displays and breeding behavior 235, feeding at grass fires 17, field characters 234, food 235, habitat and general habits 235, history and subspecies 233, measurements and weights 234, range and status 235, voice 234 White-crowned Cuckoo 424–425, plate 13 breeding 425, description 424, field characters 425, food 425, habitat and general habits 425, history 425, measurements and weights 425, range and status 425, voice 425 White-crowned Koel 424 White-eared Bronze-cuckoo 413–414, plate 12 breeding 414, description 413, field characters 414, food 414, habitat and general habits 414, measurements and weights 414, range and status 414, voice 414 White-necked Coucal 210 wild pigs and peccaries, birds following 19 wing formula xx wing length and body mass 39 wing shape 40, 41 variations 38, 39 Yellow-billed Cuckoo 339–344, plate 9 breeding 344, description 339, displays and breeding behavior 343, field characters 340, food 343, habitats and general habits 343, measurements and weights 339, night migration 25, range and status 340, tent caterpillar diet 17, voice 340 Yellow-billed Malkoha 306–308, plate 8 breeding 308, description 306, field characters 307, food 307, habitat and general habits 307, measurements 307, range and status 307, subspecies 306, voice 307 Yellow-throated Cuckoo 399–400, plate 11 breeding 400, description 399, field characters 399, food 400, habitat and general habits 400, measurements and weights 399, range and status 400, voice 400 Yellow-throated Glossy Cuckoo 399 Yellow-throated Green Cuckoo 399 Zanclostomus 293 Z. javanicus 293–294 zygodactyl feet 27, 53