<<

University of Calgary PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2016 Nesting and Incubation in : Morphological and Statistical Investigations into the Study of , , and

Tanaka, Kohei

Tanaka, K. (2016). Nesting and in Dinosaurs: Morphological and Statistical Investigations into the Study of Eggs, Eggshells, and Nests (Unpublished doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/27073 http://hdl.handle.net/11023/3505 doctoral thesis

University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

UNIVERSITY OF CALGARY

Nesting and Egg Incubation in Dinosaurs: Morphological and Statistical Investigations into the

Study of Eggs, Eggshells, and Nests

by

Kohei Tanaka

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

GRADUATE PROGRAM IN GEOLOGY AND GEOPHYSICS

CALGARY, ALBERTA

DECEMBER, 2016

© Kohei Tanaka 2016

i

Abstract (e.g., crocodylians, dinosaurs, and ) are the most diverse and successful terrestrial vertebrates. An understanding of the nesting strategies in both extinct (e.g., non-avian dinosaurs) and extant archosaurs (i.e., crocodylians and birds) is crucial for advancement of our knowledge on the and diversification of this group. However, nesting methods and behaviors of non-avian dinosaurs are still poorly understood due to the limitations of the fossil record. In this dissertation, certain features of eggs and nests in dinosaurs (e.g., clutch size, egg mass, substrates of nests, water vapor conductance of eggs, and porosity) are compared with those of their closest living relatives (i.e., birds and crocodylians) and aspects of nesting (i.e., type, incubation behavior, incubation heat source, and incubation period) are inferred and reconstructed. Findings in this dissertation suggest that nests and nesting styles among non-avian dinosaurs were diverse, and that -like traits were acquired throughout their evolution. Analyses of eggs and eggshell porosity indicate that more basal dinosaurs (i.e., ornithischians, sauropodomorphs, ) completely covered their eggs with nest materials during incubation, although more derived forms (e.g., oviraptorosaurs, troodontids) used open nests, like modern birds, in which the eggs were not fully buried. The lithologies of the clutches of basal dinosaurs reveal their nests were probably incubated with external heat sources (e.g., microbial respiration, solar radiation), like those of modern crocodylians and birds. Distribution and lithologies of some ornithischian and some sauropodomorph clutches show that heat from microbial respiration, in particular, was used for incubation, whereas other sauropodomorphs may have used inorganic heat sources, such as solar radiation. More derived dinosaurs (i.e., maniraptorans) had eggshell porosities and clutch lithologies that indicate their nests were partially open, indicating that these taxa brooded their eggs. Regardless of the type of nest, heat source, or incubation behavior, incubation period of most non-avian dinosaurs examined was relatively short, more comparable to that of birds than crocodylians. Major dinosaur (and ) show diversity in their nesting and incubation, and also reveal a transition to more bird-like nesting features through evolution.

ii

Acknowledgements I would like to thank a number of people who helped and supported my Ph.D. program. Firstly, I am significantly indebted to my supervisor, Dr. Darla Zelenitsky, who guided and encouraged me throughout the program. My Ph.D. research was further improved by her endless support and patience. Along with Dr. Zelenitsky, I would like to thank my supervisory committee, Drs. François Therrien and Jason Anderson for their significant input and direction. I am also grateful to the rest of my examiners, Drs. Jessica Theodor, Don Brinkman, Alex Dutchak, and Dan Georgescu for their significant feedback. My special thanks go to Dr. Yoshitsugu Kobayashi, who gave me endless supports and invaluable advice. I would like to thank Drs. Donald Hoyt (California State Polytechnic University, Pomona, California), Roger Seymour (University of Adelaide, Adelaide, Australia), Chris (Johns Hopkins University, Baltimore, Maryland), and Chris DeBuhr (University of Calgary) for their significant input. I am also grateful to Drs. Lars Schmitz (Claremont McKenna, Pitzer, and Scripps Colleges, Claremont, California), Jessica Theodor, Anthony Russell (University of Calgary), Nick Campione (Uppsala University, Uppsala, Sweden), Theodor Garland Jr. (University of California Riverside, Riverside, California), and David Varricchio (Montana State University, Bozeman, Montana) for their technical support. I am deeply indebted to Gregory Watkins-Colwell (Yale Peabody Museum of Natural History, New Haven, Connecticut), Kristof Zyskowski (Yale Peabody Museum of Natural History), Peter Brazaitis (Yale Peabody Museum of Natural History), Mark Peck (Royal Ontario Museum, Toronto, ), Kevin Torregrosa (St. Augustine Farm Zoological Park, St. Augustine, Florida), José Rosado (Museum of Comparative Zoology, Cambridge, Massachusetts), Tsuyoshi Takahashi (Museum of Comparative Zoology), B. J. Gill (Auckland War Memorial Museum, Auckland, New Zealand), Daisuke Suzuki (Sapporo Medical University, Sapporo, Japan), Hiroki Echizenya (Hokkaido University Museum), Dawna MacLeod (Royal Tyrrell Museum of Palaeontology, Drumheller, Canada), Brandon Strilisky (Royal Tyrrell Museum of Palaeontology), Kenneth Krysko (Florida Museum of Natural History, Gainesville, Florida), and Kent Vilet (University of Florida, Gainesville, Florida) for specimen access. Some eggshell materials, which are now registered and housed at the Yale Peabody Museum and the Hokkaido University Museum, were kindly

iii

provided by the Connecticut's Beardsley Zoo (Bridgeport, Connecticut), Hamamatsu City Zoo (Hamamatsu, Japan), Jacksonville Zoo and Gardens (Jacksonville, Florida), Kobe City Oji Zoo (Kobe, Japan), Philadelphia Zoo (Philadelphia, Pennsylvania), Sapporo City Maruyama Zoo (Sapporo, Japan), and St. Augustine Alligator Farm Zoological Park. I am also grateful to Annie Quinney, Junchang Lü, Yuong-Nam Lee, Sean Modesto, Frank Hadfield, Kirstin Brink, Caleb Brown, Matthew Ceasar, Kentaro Chiba, Julius Csotonyi, Robin Cuthbertson, Jordan Mallon, Amanda McGee, Haruo Saegusa, Tadahiro Ikeda, Huali Chang, Shaohui Fan, Songhai Jia, Hua Li, Hanyong Pu, Zhanfu Shao, Li Xu, Laiping , Hui Zhong, Eamon Drysdale, Rachel Nottrodt, Jared Voris, and colleagues at Anderson’s, Russell’s, and Theodor’s labs at the University of Calgary as well as colleagues and volunteers at Kobayashi's lab at Hokkaido University for their encouragement and scientific discussion. Finally, many thanks go to Takahiko and Makiko Tanaka, my two sisters, my niece Tsumugi, Sachiko Ohnishi, Hana, Adam, Kathy, and KJ Jansen with Chixdiggit, Vicky, Gary, Cloé, Françoise, and Jason Lawrence, Kyle Zelenitsky Therrien, Kei Nakano, Jun-Ichi Washio, and Toshikazu Toyama. My Ph.D. projects were supported by grants from the Japan Student Services Organization (JASSO), the Yoshida Scholarship Foundation, the Department of Geoscience (University of Calgary), Dinosaur Research Institute, the Society of Vertebrate Paleontology, as well as through a Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant awarded to my supervisor (Darla Zelenitsky).

iv

Dedication

For Sachiko Ohnishi

v

Table of Contents

Abstract ...... ii Acknowledgements ...... iii Dedication ...... v Table of Contents ...... vi List of Tables ...... ix List of Figures ...... xi List of Abbreviations ...... xiii

CHAPTER 1: GENERAL INTRODUCTION ...... 1

1.1 INTRODUCTORY STATEMENT ...... 1

1.2 NESTING METHODS AND BEHAVIORS IN EXTANT ARCHOSAURS ...... 2

1.3 DINOSAUR EGGS AND EGGSHELLS IN THE FOSSIL RECORD ...... 2

1.4 PREVIOUS WORK ON EGG WATER VAPOR CONDUCTANCE AND NEST TYPE IN DINOSAURS ...... 3

1.5 PREVIOUS WORK ON INCUBATION IN DINOSAURS ...... 4

1.6 OBJECTIVES AND HYPOTHESES ...... 6

CHAPTER 2: COMPARISONS BETWEEN EXPERIMENTAL AND MORPHOMETRIC WATER VAPOR CONDUCTANCE IN EGGS OF EXTANT BIRDS AND : IMPLICATIONS FOR PREDICTING NEST TYPE IN DINOSAURS ...... 12

2.1 INTRODUCTION ...... 12 2.1.1 Water vapor conductance of eggs ...... 13

2.2 MATERIALS AND METHODS ...... 14 2.2.1 Selection of taxa ...... 14 2.2.2 Experimental water vapor conductance (Gex) ...... 15 2.2.3 Morphometric water vapor conductance (Gmo) ...... 15 2.2.4 Egg mass ...... 17 2.2.5 Statistical analyses ...... 18 2.2.5.1 Conventional statistical approach ...... 18 2.2.5.2 Phylogenetically corrected statistical approach ...... 19

2.3 RESULTS ...... 20

vi

2.3.1 Descriptive statistics ...... 20 2.3.2 Bivariate plots of GH2O ...... 20 2.3.3 Passing–Bablok linear regression ...... 21 2.3.4 Bland–Altman plot ...... 21

2.4 DISCUSSION ...... 22

CHAPTER 3: EGGSHELL POROSITY PROVIDES INSIGHT ON EVOLUTION OF NESTING IN DINOSAURS ...... 36

3.1 INTRODUCTION ...... 36

3.2 MATERIALS AND METHODS ...... 38 3.2.1 Relationship between water vapor conductance and eggshell porosity ...... 38 3.2.2 Selection of extant taxa ...... 39 3.2.3 Nest classification for extant taxa ...... 40 3.2.4 Selection of fossil eggs/ootaxa ...... 40 3.2.5 Eggshell porosity ...... 42 3.2.6 Egg mass ...... 44 3.2.7 Phylogenetic distribution of nest type ...... 44 3.2.8 Analysis of covariance ...... 45 3.2.9 Discriminant analysis ...... 46

3.3 RESULTS ...... 47 3.3.1 Estimated D statistic ...... 47 3.3.2 Analysis of covariance (ANCOVA) ...... 48 3.3.3 Discriminant analysis ...... 48

3.4 INTERPRETATION OF STATISTICAL RESULTS ...... 50

3.5 DISCUSSION ...... 51

CHAPTER 4: NEST MATERIAL, NEST TYPE, AND INCUBATION HEAT SOURCES IN DINOSAURS: IMPLICATIONS FOR DINOSAUR NESTING AT HIGH LATITUDE ...... 76

4.1 INTRODUCTION ...... 76

4.2 RESEARCH APPROACHES ...... 77

4.3 MATERIAL AND METHODS ...... 78 4.3.1 Datasets and statistical tests for extant taxa ...... 78 4.3.2 Datasets and statistical tests for fossil taxa/ootaxa ...... 79

4.4 RESULTS ...... 80 vii

4.4.1 Dataset for extant taxa ...... 80 4.4.2 Dataset for dinosaur taxa/ootaxa ...... 81

4.5 DISCUSSION ...... 82 4.5.1 Nesting methods of extant species ...... 83 4.5.2 Nest structures, nesting substrates, and incubation heat sources of extinct dinosaurs ...... 85 4.5.3 Distribution of dinosaur eggs and implications for dinosaur nesting at high latitude...... 86

CHAPTER 5: ESTIMATION OF INCUBATION PERIOD IN DINOSAURS ...... 106

5.1 INTRODUCTION ...... 106

5.2 MATERIALS AND METHODS ...... 107 5.2.1 Selection of living taxa ...... 107 5.2.2 Selection of fossil taxa/ootaxa ...... 108 5.2.3 Values for incubation period ...... 108 5.2.4 Selection of independent variables ...... 109 5.2.5 Values for egg mass, eggshell porosity, and pore density ...... 111 5.2.6 Nest type ...... 111 5.2.7 Statistical analyses ...... 111

5.3 RESULTS ...... 113

5.4 DISCUSSION ...... 115

CHAPTER 6: GENERAL DISCUSSION ...... 127

CHAPTER 7: CONCLUSIONS ...... 130

APPENDICES ...... 133

REFERENCES ...... 364

viii

List of Tables

Table 1.1. List of dinosaur oofamilies and possible parental taxa...... 8

Table 2.1. Comparisons of two methods for estimating water vapor conductance of archosaur eggs...... 25

Table 2.2. Lists of variables and equations used for estimation of eggshell porosity (variables and definitions from Deeming, 2006 and references therein)...... 26

Table 2.3. Results of descriptive statistics for GH2O datasets...... 27

Table 2.4. Results of regression analyses by the conventional model, phylogenetically corrected model, and Passing–Bablok method...... 28

Table 2.5. Results of the Bland–Altman plots between differences of the experimental (log Gex) and morphometric (log Gmo) methods and their mean values...... 30

Table 3.1. List of extinct archosaur taxa/ootaxa with estimated egg mass (M) and eggshell −1 porosity (Ap∙Ls ) used in this study...... 55

Table 3.2. List of variables used for this study, modified from Tanaka and Zelenitsky (2014a)...... 59

Table 3.3. List of equations used for this study, modified from Tanaka and Zelenitsky (2014a)...... 60

Table 3.4. Results of conventional OLS regression models for living archosaur species...... 61

Table 3.5. Results of conventional and phylogenetically-corrected ANCOVA for living archosaur species...... 62

Table 3.6. Cross-classification/confusion matrix from LDA and pFDA...... 63

Table 3.7. Inferred nest types for extinct archosaurs based on the linear discriminant analysis. . 64

Table 3.8. Inferred dinosaur nest types based on the phylogenetic flexible discriminant analysis...... 67

Table 4.1. List of clutch specimens observed at the Royal Tyrrell Museum of Palaeontology (TMP), Drumheller, Alberta, Canada...... 89

Table 4.2. List of ootaxa and their possible parental taxa used in this study...... 90

ix

Table 4.3. Results of one-way χ2-tests for the in-situ dataset of dinosaur ootaxa...... 91

Table 4.4. Results of one-way χ2-tests for the ex-situ dataset of dinosaur ootaxa...... 92

Table 4.5. Number of occurrences of major dinosaur clutches among countries...... 93

Table 4.6. Ranges of absolute paleolatitudes in major dinosaur ootaxa...... 94

Table 5.1. Results of simple linear regression analyses in extant archosaur species...... 119

Table 5.2. Results of multiple regression analyses in extant archosaur species...... 120

Table 5.3. Estimates of incubation period ± prediction errors (days) in dinosaurs based on multiple regression models...... 121

x

List of Figures

Figure 2.1. Porosity of archosaur eggshell...... 31

Figure 2.2. Bivariate plots of experimental (log Gex) and morphometric (log Gmo) conductance values showing positive correlations between these variables: (a) all species, (b) , (c) , and (d) ...... 32

Figure 2.3. Bivariate plots of water vapor conductance (log GH2O) and egg mass (log M) showing positive correlations between these variables: (a) all species, (b) Anseriformes, (c) Charadriiformes, and (d) Galliformes...... 33

Figure 2.4. Passing–Bablok regression plots for the datasets of experimental (log Gex) and morphometric (log Gmo) conductance values in living archosaurs: (a) all species, (b) Anseriformes, (c) Charadriiformes, and (d) Galliformes...... 34

Figure 2.5. Bland–Altman plot of differences and mean values between experimental (log Gex) and morphometric (log Gmo) conductance values: (a) all species, (b) Anseriformes, (c) Charadriiformes, and (d) Galliformes...... 35

Figure 3.1. Porosity of archosaur eggshell...... 68

Figure 3.2. Bivariate plot of eggshell porosity and egg mass between living covered and open nesters...... 69

Figure 3.3. Bivariate plot of eggshell porosity and egg mass in both living and extinct archosaur taxa/ootaxa...... 70

Figure 3.4. Misclassification rate of pFDA for living species through changing Pagel's lambda values...... 71

Figure 3.5. Comparison of the discriminant function between covered and open nesters in living and fossil archosaurs...... 72

Figure 3.6. Inferred nest type for six extinct archosaurs as a function of Pagel's lambda values...... 73

Figure 3.7. Evolution of nest types among archosaurs...... 74

Figure 4.1. Schematic diagram of tests and comparisons conducted in this study...... 95

Figure 4.2. Boxplot of mean nest (Tnest) and ambient air temperature (Tair) among covered nests that use heat from microbial respiration, geothermal energy, and solar radiation...... 96

xi

Figure 4.3. Comparisons of nest materials/substrates between extant covered nests that use inorganic heat sources and organic (and some inorganic) heat sources: (a) barplots and (b) pie charts...... 97

Figure 4.4. Comparisons of nest materials/substrates between extant mound and in-filled hole nests: (a) barplots and (b) pie charts...... 98

Figure 4.5. Comparisons of nest lithology among dinosaur oofamilies: (a) in-situ dataset and (b) ex-situ dataset...... 99

Figure 4.6. Ratio of or pedogenetic features found in dinosaur nests...... 100

Figure 4.7. Boxplot of absolute paleolatitude in /eggshell localities...... 101

Figure 4.8. Distributions of current covered nesters (i.e., crocodylians and ): (a) mound nesters that use heat from microbial respiration, (b) in-filled hole nesters that use heat from geothermal energy, and (c) in-filled hole nesters that use heat from solar radiation...... 102

Figure 4.9. Inferred nest structures and incubation heat sources for dinosaurs...... 103

Figure 4.10. (a) and (b) Early localities of dinosaur egg/eggshell remains...... 104

Figure 4.11. localities (dots) of dinosaur egg/eggshell remains: (a) hadrosaurs (Spheroolithidae) and sauropods laying megaloolithids, (b) sauropods laying faveoloolithids, and (c) Dendroolithidae, oviraptorosaurs () and troodontids ()...... 105

Figure 5.1. Bivariate plots between incubation period and predictor variables (i.e., egg mass, eggshell porosity, and pore density)...... 122

Figure 5.2. Comparison of percent prediction error (%) of incubation period estimates among regression models...... 123

Figure 5.3. Bivariate plot of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa...... 124

Figure 5.4. Comparisons of predictor variables in living and extinct archosaur taxa/ootaxa, including some outlier taxa of living bird groups...... 125

Figure 5.5. Transition of incubation period in archosaurs...... 126

Figure 6.1. Transition of reproductive traits in archosaurs...... 129

xii

List of Abbreviations

Institutional Abbreviations

AIM Auckland Institute and Museum, Auckland, New Zealand CM Carnegie Museum of Natural History, Pittsburgh, Pennsylvania HEC Hirsch Egg Catalogue, University of Colorado Museum, Boulder, Colorado MCZ Museum of Comparative Zoology, Cambridge, Massachusetts ROM Royal Ontario Museum, Toronto, Canada TMP Royal Tyrrell Museum of Palaeontology, Drumheller, Canada UHR Hokkaido University Museum, Sapporo, Japan YPM R. Herpetology Collection at the Yale Peabody Museum, New Haven, Connecticut ZEC Zelenitsky Egg Catalogue, the University of Calgary, Calgary, Canada

Technical Abbreviations

A Mean individual pore area (mm2) AIC Akaike Information Criterion ANCOVA Analysis of covariance 2 Ap Total pore area in an egg (mm ) −1 Ap∙Ls Eggshell porosity (mm) 2 As Eggshell surface area (mm ) B Maximum egg breadth (mm) 9 −1 −1 c Unit conversion constant (1.56·10 mg H2O∙s∙day ∙mol ) C Clay CI Confidence interval D Pore density (number of pores∙mm−2) D Dendroolithidae Mean difference ΔP Gradient of water vapor pressure between the inside of the egg and the nest (Torr) d.f. Degree of freedom 2 −1 DH2O Diffusion coefficient of water vapor (mm ∙s ) E Elongatoolithidae F Test statistic for ANCOVA F Fraction of water vapor lost during incubation, relative to egg mass Fa Faveoloolithidae −1 −1 Gex Experimental water vapor conductance (mg H2O∙day ∙Torr ) −1 −1 GH2O Water vapor conductance (mg H2O∙day ∙Torr ) −1 −1 Gmo Morphometric water vapor conductance (mg H2O∙day ∙Torr ) H In-filled hole nest I Inorganic heat source xiii

Ip Incubation period (day) L Maximum egg length (mm) L. Linnaeus LDA Linear discriminant analysis log Logarithm Ls Eggshell thickness (i.e., pore length) (mm) M Egg mass (g) M Mound nest Me −1 MH2O Rate of water vapor diffusion (i.e., the daily loss of water vapor) (mg H2O∙day ) n Sample size N Total number of pores in the egg N North NA Not available or not applicable NT Nest type O Organic heat source (with possibly some inorganic heat) OLS Ordinary least-squares p Probability for a hypothesis test PB Passing-Bablok regression PcANCOVA Phylogenetically-corrected ANCOVA Pegg Absolute water vapor pressure of the inside of the egg (Torr) pFDA Phylogenetic flexible discriminant analysis PGLS Phylogenetically-generalized least-squares assumed Brownian motion process Pnest Absolute water vapor pressure of the nest (Torr) PPE Percent prediction error (%) Pre Preprismatoolithus Pri Prismatoolithidae PS Plant material and/or soil R Universal gas constant (6.24·107 mm3∙Torr∙mol−1∙°K−1) r Pearson’s correlation coefficient r2 Coefficient of determination RegOU Phylogenetic regression with Ornstein-Uhlenbeck process S South Sa SD Standard deviation SE Standard error SEE Standard error of the estimate Sp Spheroolithidae t Test statistic for T-test T Absolute temperature of incubation (°K) Tair Mean ambient air temperature of the nesting habitat (°C) Tnest Mean nest temperature (°C) 3 V Egg volume (mm ) VIF Variance inflation factor xiv

CHAPTER 1: GENERAL INTRODUCTION

1.1 INTRODUCTORY STATEMENT Archosaurs (e.g., crocodylians, dinosaurs, and birds) are the most diverse and successful terrestrial vertebrates. An understanding of the nesting strategies in both extinct (e.g., non-avian dinosaurs) and extant archosaurs (i.e., crocodylians and birds) is crucial for advancement of our knowledge of the evolution and diversification of this group. Although nesting methods and behaviors are observable in extant archosaurs (e.g., Coombs, 1989; Deeming, 2002a), those of non-avian dinosaurs have been inferred (e.g., Seymour, 1979; Varricchio et al., 1997, 2008; Grellet-Tinner and Fiorelli, 2010; Ruxton et al., 2014; Wiemann et al., 2015) but are still poorly understood due to the limitations of the fossil record. However, aspects of nesting in extinct dinosaurs can be inferred or reconstructed by comparisons of their eggs and nests to those of their closest living relatives (i.e., birds and crocodylians). It has been shown that certain features of extant archosaur eggs and nests – specifically egg mass, clutch size (i.e., number of eggs in a nest), water vapor conductance of eggs (i.e., diffusive capacity of the eggshell), eggshell porosity (i.e., total pore area in an egg divided by pore length), and materials/substrates of nests – are related to aspects of their incubation, including nest type (e.g., Ar and Rahn, 1985), incubation behavior (e.g., brooding) (e.g., Seymour and Ackerman, 1980), incubation heat source (e.g., Ferguson, 1985; Magnusson et al., 1985; Booth and Jones, 2002), incubation period (i.e., the time interval from oviposition to hatching) (e.g., Rahn and Ar, 1974; Deeming et al., 2006), metabolism of (e.g., Hoyt and Rahn, 1980), or developmental maturity at hatching (i.e., precocial vs. altricial) (e.g., Birchard et al., 2013). Not all of these traits are observable in the fossil record; however, clutch size, egg mass, lithology of nests, water vapor conductance of eggs, and eggshell porosity can be observed or estimated from fossil materials, and such traits can then be compared to those of extant species. The research conducted here aims to shed light on the diversity and evolution of nesting strategies among archosaurs. This will be achieved by compiling and comparing large datasets of features/traits related to egg incubation (e.g., nest types, incubation behaviors, clutch size, nest materials/substrates, incubation heat source, nest temperature, incubation period, egg mass, water vapor conductance of eggs, and eggshell porosity) in extant archosaurs. These comparisons will be used for developing new methods for predicting nest types, incubation behaviors, incubation heat

1 sources, and incubation period for archosaurs, which can be used in turn to predict/infer incubation traits in non-avian dinosaurs that are not preserved directly in the fossil record.

1.2 NESTING METHODS AND BEHAVIORS IN EXTANT ARCHOSAURS Extant crocodylians and birds generally differ in their nest types and nesting strategies. Extant crocodylians build covered nests, where eggs are completely covered with nest materials/substrates (e.g., sand, soil, plant materials: Greer, 1970; Campbell, 1972), and lay relatively small eggs (only 0.1–1.2% of adult body masses: calculated from the dataset by Werner and Griebeler, 2013) with large clutch sizes (on average 12–55 eggs: Ferguson, 1985; Thorbjarnarson, 1996). Compared with crocodylians, avian clutch sizes are generally small (1–20 eggs: Gill, 2007) and their incubation period tends to be much shorter on average (10–90 days: Rahn and Ar, 1974), although relative egg size to the adult body mass tends to be larger than that of crocodylians (in most cases 2–11%: Gill, 2007). Crocodylian covered nests include mound (i.e., pile of nest materials on the ground) and in-filled hole nests (i.e., underground burrow) (Greer, 1970; Campbell, 1972). Incubation takes an average 68–99 days (Birchard and Marcellini, 1996), using incubation heat from surrounding environments (e.g., heat from solar radiation and decaying plant materials via microbial respiration: Ferguson, 1985; Magnusson et al., 1985). Among extant birds, such incubation style (i.e., mound and in-filled hole nests using heat from solar radiation, geothermal activities such as volcanoes and hot springs, and microbial respiration) is only found in family Megapodiidae (order Galliformes), which independently acquired this incubation style (Dekker and Brom, 1992; Harris et al., 2014). Eggs of the remaining extant bird species are incubated by adult contact incubation (i.e., brooding) in open nests, where eggs are at least partially exposed (Deeming, 2002a). Open nests are composed of various materials (e.g., plants, lichen, fungi, , fur, silk, stones, or mud: Hansell, 2000) and the structure of open nests is highly variable, from a simple scrape on the ground to a sophisticated domed nest structure on a tree (Hansell and Deeming, 2002).

1.3 DINOSAUR EGGS AND EGGSHELLS IN THE FOSSIL RECORD In non-avian dinosaurs, the size and morphology of the clutches, eggs, and eggshells are highly diverse. The eggs were laid in clusters of up to 40 eggs (e.g., Young, 1965; Zhao et al., 1991; Wang and Zhou, 1995; Mateus et al., 1997; Varricchio et al., 1997; Horner, 1999; Vila et al.,

2

2010b). These clusters generally contain one or more layers of eggs that are either randomly arranged (e.g., Vila et al., 2010b) or arranged in a certain pattern (e.g., circular: Young, 1954, 1965). Dinosaur eggs are spherical to elongate in shape, including asymmetric oval like those of extant birds (Deeming and Ruta, 2014). Egg masses are estimated to have ranged from approximately 100 g for the smallest to up to 7 kg for the largest (Tanaka et al., 2016). Their eggshell is a rigid calcitic structure like those of birds and crocodiles, which is pierced by numerous pores. These pores vary in shape from straight and tubular to irregular and branched (Erben, 1979; Grellet-Tinner et al., 2012; Hechenleitner et al., 2016b). The eggshell is made up of shell units that consist of various structural layers, which exhibit various ultrastructures (e.g., crystal morphology: Erben, 1970; Mikhailov, 1991). Fossil eggs and eggshells are usually not found in close association with skeletal remains (e.g., embryos, hatchlings, gravid adults, and adults in nests), and thus precise taxonomic affinities of most dinosaur eggs are unknown. Because of this, fossil eggs and eggshells are classified using a parataxonomic scheme (i.e., egg parataxonomy or ootaxonomy) that is independent from the based on skeletal remains (Mikhailov et al., 1996; Zelenitsky and Hirsch, 1997). Features such as egg size, shape, shell thickness, outer surface ornamentation, and eggshell microstructures are used for the parataxonomic classification. So far, 19 dinosaur oofamilies (i.e., egg families), including over 60 oogenera (i.e., egg genera) and over 150 oospecies (i.e., egg species), have been identified (Table 1.1). Only a single oospecies has been ascribed to a precise dinosaur species (i.e., levis for an egg of fomosus: Varricchio et al., 2002), although some oogenera and oofamilies have been ascribed to more inclusive dinosaur taxa (Table 1.1). The high diversity of dinosaur ootaxa suggests that the eggs of dinosaurs are highly variable. Some of the variability likely reflects different nesting methods and behaviors.

1.4 PREVIOUS WORK ON EGG WATER VAPOR CONDUCTANCE AND NEST TYPE IN DINOSAURS Regardless of the differences in incubation methods (e.g., open vs. covered nests), all extant archosaur species lay calcified, rigid eggshell that is pierced by numerous pores, which are the pathways for gas exchange (i.e., embryonic respiration and water vapor diffusion). During incubation, archosaur eggs lose 10–23% of their water (Ar and Rahn, 1980; Lutz and Dunbar-Cooper, 1984) due to the gradient of water vapor pressure between the inside and the

3 outside of the eggs (Ar et al., 1974). The diffusive capacity of eggshell is referred to as the water vapor conductance of eggshell (Ar et al., 1974). Because of the differences of size, shape, and number of pore canals, values of water vapor conductance vary among species (e.g., Ar and Rahn, 1985).

Water vapor conductance (GH2O) has been used to investigate physiological properties of the eggs in extant archosaurs (e.g., Ar et al., 1974; Ar and Rahn, 1985; Clark et al., 2010; Jaeckle et al., 2012) and as a proxy for nest type (i.e., open vs. covered) in non-avian dinosaurs (e.g.,

Seymour, 1979; Deeming, 2006; Varricchio et al., 2013). Nest type, inferred from GH2O, provides significant insight into incubation behavior in archosaurs because nest type is strongly correlated with incubation behavior in that all eggs laid in uncovered, open nests are incubated via brooding (i.e., adult contact incubation), whereas eggs in covered nests are always incubated with external heat sources (e.g., Deeming, 2002a). In extant species, GH2O has been estimated using an experimental method (Gex) initially developed by Ar et al. (1974); however, this method requires fresh eggs and is not appropriate for use with extinct taxa. An alternative approach to estimate

GH2O in extant species is a theoretical formula derived from Fick's first law of gas diffusion, which uses morphometric data of the pore canals of eggshell (Ar et al., 1974), here referred to as morphometric GH2O (Gmo). This method was subsequently applied to non-avian dinosaurs by researchers investigating their nest types (e.g., Seymour, 1979; Deeming, 2006; Varricchio et al.,

2013), following the idea that Gex reflects nest type of extant taxa (Seymour, 1979). However, it is not known if the Gex dataset is directly comparable to the Gmo estimations for extant taxa as the morphometric data are limited. As a result, the use of Gmo data to infer nest type in non-avian dinosaurs remains questionable as morphometric data in extant taxa have never been related to nest type or incubation behavior. If the two methods can be shown to produce comparable estimates of GH2O, then it is potentially valid to use the morphometric GH2O method to infer nest type in dinosaurs. Alternatively, if the methods are not in agreement, it may be necessary to develop an alternative method to infer nest type of extinct archosaurs.

1.5 PREVIOUS WORK ON INCUBATION IN DINOSAURS Very little is known about incubation in non-avian dinosaurs other than the incubation behavior (i.e., brooding vs. non-brooding), which has been inferred from nest type based on GH2O estimates (e.g., Seymour, 1979; Deeming, 2006; Varricchio et al., 2013). Incubation styles and

4 behaviors, however, are well known in many extant species, and these can be related to certain features of eggs and nests (e.g., egg mass, eggshell porosity, and nest materials/substrates: Ar et al., 1974; Rahn and Paganelli, 1990; Deeming, 2002b; Somaweera and Shine, 2013). For example, incubation behavior is strongly related to nest type among extant species as eggs in open nests are incubated via brooding, whereas those in covered nests are incubated with external heat sources

(e.g., Deeming, 2002a); thus nest type, inferred from GH2O estimates, in non-avian dinosaurs has been used as a basis to predict their incubation behavior (e.g., Seymour, 1979; Deeming, 2006; Varricchio et al., 2013). While nest type and associated incubation behavior in non-avian dinosaurs have received the most study, other aspects of incubation, such as incubation heat source, is poorly understood in non-avian dinosaurs. Research on incubation heat source is important to understand the paleogeography of dinosaur nesting sites because the availability of external incubation heat sources (i.e., solar radiation, geothermal activities, and microbial respiration) is related to climate and geographic distribution of nesting sites. Incubation heat sources thus could be a factor that affects the distribution of dinosaur nesting sites, especially for species that do not brood their eggs. Although there is limited research on incubation heat sources for dinosaurs, their eggs and eggshells have been reported from various regions (e.g., Asia, , , North America, and South America: Carpenter, 1999), depositional environments (e.g., fluvial, lacustrine, littoral, and eolian: Paik et al., 2012), and sediments (e.g., siliciclastic , siltstone, , and as well as carbonate rocks, with or without pedogenic features: e.g., Khosla and Sahni, 1995; Varricchio et al., 1999; Liang et al., 2009). In particular, evidence of dinosaur eggshells and babies in high Arctic (e.g., Godefroit et al., 2009; Fiorillo et al., 2014) raises an intriguing question as to how non-avian dinosaurs acquired incubation heat and incubated their eggs in cooler environments. Although adult contact incubation (i.e., brooding) transfers body heat to the eggs in open nesters, eggs in covered nests require external heat sources for their incubation. Because extant covered nesters (i.e., crocodylians and megapodes) generate and/or transfer incubation heat (Ferguson, 1985; Magnusson et al., 1985; Booth and Jones, 2002) with nest materials/substrates, certain nest materials/substrates are likely associated with certain incubation heat sources. Thus, nest materials/substrates (inferred from the lithology of nests) in non-avian dinosaurs can potentially be used to predict their incubation heat sources. Incubation period, measured as the time interval from oviposition to hatching, is not

5 directly measurable in non-avian dinosaurs. Although it has been shown that there is a positive correlation between egg mass and incubation period in extant birds (e.g., Rahn and Ar, 1974; Ar and Rahn, 1978; Deeming et al., 2006), this may not be an accurate method based on their data because incubation period is not only affected by egg mass (Ricklefs and Stark, 1998; Martin,

2002; Deeming et al., 2006) but also by other factors such as clutch mass, egg temperature, GH2O, growth rate and metabolism of , parental behavior, nest location and structure, climatic conditions, and phylogeny (Deeming et al., 2006). A statistically-reliable regression model has yet to be developed for predicting incubation period in extant archosaurs, although one study has predicted incubation period of Sauropodomorpha based on the allometry of egg mass and incubation period using extant archosaurs (Ruxton et al., 2014). Since various features affect incubation period (e.g., Deeming et al., 2006; Zimmermann and Hipfner, 2007), these factors could potentially be used to predict incubation period in extant species as well as in non-avian dinosaurs.

1.6 OBJECTIVES AND HYPOTHESES The main objective of this research is to predict nest type, incubation behavior, incubation heat source, and incubation period of non-avian dinosaurs in order to elucidate the transitions of nesting strategies/behavior through archosaur evolution. To achieve this, the evolutionary history of these incubation traits (i.e., nest type, incubation behavior, incubation heat source, and incubation period) will be documented using large datasets of over 150 species of extant and extinct species. These traits related to incubation will be examined with the following hypotheses. Initially, it is explored if the previous method to predict nest types of non-avian dinosaurs is statistically valid by testing agreement between datasets for Gex and Gmo in living species (Chapter 2). If the previous method is not appropriate, then a new approach will be developed for predicting nest types in dinosaurs. Eggshell porosity could be an alternative to GH2O because the method used for estimation of eggshell porosity is consistent between fossil and fresh eggs. Thus, subsequently, it will be tested if there is significant difference in eggshell porosity among nest types/incubation behaviors (Chapter 3). If a significant difference is revealed, then eggshell porosity will be used for predicting nest type and associated incubation behavior (e.g., brooding) among non-avian dinosaurs. In Chapter 4, it is explored if incubation heat sources of extant archosaur nests are statistically correlated to nest materials/substrates (e.g., sand, soil, and plant

6 materials). If there are correlations, then incubation heat source could be inferred based on the lithology of nests (potentially corresponding to nest materials/substrates) for non-avian dinosaurs, which allows discussion of the paleogeography of dinosaur nesting sites. Finally, I test if incubation period is correlated with parameters that are related to incubation (e.g., egg mass, eggshell porosity, and nest type) in extant archosaurs (Chapter 5). If there are correlations between incubation period and these parameters in extant species, then these parameters will be used for predicting incubation period of non-avian dinosaurs. The research, including published projects (Chapter 2 as Tanaka and Zelenitsky, 2014a and Chapter 3 as Tanaka et al., 2015), was conducted by the first author (Kohei Tanaka) (e.g., research design, data collection, data analyses, and manuscript writing), overseen by my supervisor (Dr. Darla Zelenitsky) and Dr. François Therrien.

7

Table 1.1. List of dinosaur oofamilies and possible parental taxa. Note that: *Dughioolithus is a synonym of Cairanoolithus (Sellés and Galobart, 2016); †Longiteresoolithus is a synonym of (Wang et al., 2010b); ‡Fusioolithus berthei was included in Fusioolithus whereas other oospecies were considered as ; §see Mohabey (1998) and Vianey-Liaud et al. (2003) for synonyms within Megaloolithus. Oofamily Number of Number of Possible parental taxa Identification based References for oogenera oospecies on: identification Arriagadoolithidae Agnolin 2 2 Alvarezsaurid? Bones associated with Agnolin et al. (2012) et al. (2012) eggs and eggshells Belonoolithidae Jackson and 1 1 ? Eggshell morphology Jackson and Varricchio Varricchio (2016) (2016) Cairanoolithidae Sellés and 1* 2 Ornithischian? Eggshell morphology Sellés and Galobart Galobart (2016) (2016) Dendroolithidae Zhao and Li 3 15 and Embryos Manning et al. (1997); (1988) (also known as possibly Kundrat et al. (2008); Phaceloolithidae: Zeng therizinosauroids Araujo et al. (2013); and Zhang, 1979) Rebeiro et al. (2014) Dictyoolithidae Zhao (1994) 4 7 Theropoda? Eggshell morphology Jin et al. (2010) Elongatoolithidae Zhao 11† 28 Embryos, adults atop Dong and Currie (1975) clutches, and a gravid (1996); Clark et al. adult skeleton (1999); Norell et al. (2001); Sato et al. (2005); Cheng et al.

8

Oofamily Number of Number of Possible parental taxa Identification based References for oogenera oospecies on: identification (2008); Weishampel et al. (2008); Fanti et al. (2012); Wang et al. (2016) Faveoloolithidae Zhao and 4 8 Sauropodomorpha? Eggshell morphology Grellet-Tinner and Ding (1976) Fiorelli (2010); Grellet-Tinner et al. (2012) Fusioolithidae Fernandez 1 1‡ Sauropodomorpha? Eggshell morphology Fernandez and Khosla and Khosla (2015) (2014) Megaloolithidae Zhao 2 22§ Titanosaurs Embryos Chiappe et al. (1998, (1979b) 2001) Montanoolithidae Zelenitsky 1 1 Non-avian Eggshell morphology Zelenitsky and Therrien and Therrien (2008b) maniraptoran (2008b)

Ovaloolithidae Mikhailov 1 8 Unknown NA NA (1991) Pinnatoolithae Fang et al. 2 6 Unknown NA NA (2009) Polyclonoolithidae Xie et al. 1 1 Unknown NA NA

9

Oofamily Number of Number of Possible parental taxa Identification based References for oogenera oospecies on: identification (2016) Prismatoolithidae Hirsch 6 18 Troodontids and Embryos, adult bones Mateus et al. (1997); (1994) non-oviraptorosaur associated with eggs, Varricchio et al. (1997, maniraptorans and eggshell 2002); Zelenitsky and (Prismatoolithus and morphology Therrien (2008a); ) and Lopez-Martinez and and Vicens (2012); Carrano Lourinhanosaurus et al. (2013) (Preprismatoolithus) Similifaveoolithidae Wang 1 2 Unknown NA NA et al. (2011) Spheroolithidae Zhao 4 15 Hadrosaurs, including Embryos Horner and Makela (1979b) and (1979); Hirsch and Hypacrosaurus Quinn (1990); Horner (1999, 2000) Stalicoolithidae Wang et al. 2 2 Unknown NA NA (2012) Tubercuoolithidae Jackson 2 2 Theropoda? Eggshell morphology Jackson and Varricchio and Varricchio (2016) (2010, 2016) Youngoolithidae Zhang 1 2 Unknown NA NA

10

Oofamily Number of Number of Possible parental taxa Identification based References for oogenera oospecies on: identification (2010) Oofamily incertae sedis 12 12 NA NA NA

11

CHAPTER 2: COMPARISONS BETWEEN EXPERIMENTAL AND MORPHOMETRIC WATER VAPOR CONDUCTANCE IN EGGS OF EXTANT BIRDS AND CROCODILES: IMPLICATIONS FOR PREDICTING NEST TYPE IN DINOSAURS

2.1 INTRODUCTION

Water vapor conductance, or diffusive capacity of the eggshell (GH2O: mg −1 −1 H2O∙day ∙Torr ), has been used to investigate physiological properties of the eggs in living archosaurs (e.g., Ar et al., 1974; Ar and Rahn, 1985; Clark et al., 2010; Jaeckle et al., 2012) and as a proxy for nest environment (i.e., nest type) in extinct taxa (e.g., Seymour, 1979; Deeming,

2006; Varricchio et al., 2013). In living species, GH2O has been estimated by an experimental (gravimetric) method initially developed by Ar et al. (1974); however, this method requires fresh eggs and thus is not appropriate for use with fossil taxa. An alternative approach to estimate

GH2O across eggshells of living species uses a theoretical formula derived from Fick’s first law of gas diffusion, which is based on morphometric data (e.g., individual pore area, pore density, and pore length) from the pore canals (Ar et al., 1974). This morphometric method was also applied to fossil eggs by researchers investigating nest types (e.g., covered and open nests) of extinct taxa (Seymour, 1979; Williams et al., 1984; Sabath, 1991; Mou, 1992; Grigorescu et al., 1994; Mikhailov et al., 1994; Sahni et al., 1994; Antunes et al., 1998; Deeming, 2006; Jackson et al., 2008; Donaire and Lopez-Martinez, 2009; Grellet-Tinner, et al. 2012; Varricchio et al., 2013;

Zhao et al., 2013). Because GH2O values based on the morphometric method have never been directly compared with nest types in living species, these researchers assumed based on prior work (e.g., Ar et al., 1974; Hoyt et al., 1979) that values from the morphometric and experimental GH2O were essentially equivalent and also followed the idea that experimental GH2O reflects nest type of living species (e.g., Seymour, 1979). However, the degree to which the morphometric values compare with the experimental values for given species is unknown because morphometric data are limited for living species. Thus, the use of morphometric GH2O data to infer nest environment in extinct taxa is questionable. Here I statistically examine if the morphometrically derived values for GH2O in living species are comparable with the experimental GH2O values by significantly increasing the dataset of living species for which

12 morphometric data are available. This study will therefore determine if it is appropriate to infer nest type of extinct taxa via comparison of morphometric GH2O of extinct taxa with experimental

GH2O of living species.

2.1.1 Water vapor conductance of eggs Water vapor conductance for the eggs of living archosaur species has been estimated by numerous studies using an experimental method. To date, GH2O values of approximately 290 species of living birds and crocodilians have been reported (Table 2.1). Water vapor conductance is defined as the rate of water vapor diffusion per unit differential partial pressure of water vapor across the eggshell:

M H 2O [2-1] GH 2O  Pegg  Pnest where MH2O is the rate of water vapor diffusion (i.e., the daily loss of water vapor: mg −1 H2O∙day ) and Pegg and Pnest, measured in Torr, are the absolute water vapor pressure of the egg interior and nest, respectively (Ar et al., 1974). For this method, herein referred to as the experimental method, fresh eggs are placed in laboratory desiccators where the ambient humidity

(equivalent to Pnest) is 0 Torr, and, at known intervals, the eggs are removed from the desiccators and weighed. Assuming that mass loss of each egg is due solely to the loss of water vapor and that Pegg is saturated at the temperature of the desiccators (Ar et al., 1974), GH2O is obtained by dividing the daily rate of mass loss of the eggs by the saturation water vapor pressure at the ambient temperature of the desiccators (see also the Materials and Methods in this chapter). Ar et al. (1974) proposed that because the pore canals of the eggshell are the only pathways for water vapor, GH2O values for eggs also can be estimated theoretically from measurements of the pore canals (i.e., total pore area in an egg and pore length or eggshell thickness) (Table 2.1). This approach is herein referred to as the morphometric method. Based on the Fick’s first law of gas diffusion, GH2O has been estimated using the following formula:

c  DH 2O Ap [2-2] GH 2O   R T Ls 9 −1 −1 where c is a unit conversion constant (1.56·10 mg H2O∙s∙day ∙mol ), DH2O is the diffusion coefficient of water vapor (mm2∙s−1) in air temperature, R is the universal gas constant (6.24·107 3 −1 −1 mm ∙Torr∙mol ∙°K ), T is the absolute temperature of incubation (°K), Ap is a total pore area in

13

2 an egg (mm ), and Ls is eggshell thickness (i.e., pore length: mm) (Ar et al., 1974; Seymour, 1979).

Although researchers investigating fossil eggs widely accept the assumption that GH2O values derived from experimental and morphometric methods are comparable for the eggs of living species (e.g., Ar et al., 1974; Hoyt et al., 1979), there has been little evidence to support this assumption. One study found that morphometric GH2O values of 21 species of Anseriformes are, on average, 10% lower than experimental GH2O values, suggesting that the two methods are generally in agreement (Hoyt et al., 1979). The dataset, however, was restricted to small sample size (n = 21) with a narrow range of egg sizes within one avian order (Anseriformes), and no statistical tests, neither conventional nor phylogenetically corrected, were conducted. Thus, a larger dataset of morphometric GH2O that includes a diverse number of living archosaur species is required to test the agreement between the experimental and the morphometric methods and to determine if comparisons between experimental GH2O of living species and morphometric GH2O of fossil eggs are valid.

2.2 MATERIALS AND METHODS

Values of GH2O were compiled for the eggs of living archosaur species from both experimental and morphometric methods to test whether outcomes of these two methods are correlated and in agreement. Values of GH2O estimated by both methods were statistically compared among all species and among orders for which enough data were available (i.e., Anseriformes, Charadriiformes, and Galliformes). Statistical comparisons were done following the techniques described by Altman and Bland (1983) and Jensen and Kjelgaard-Hansen (2006).

2.2.1 Selection of taxa

Experimental and morphometric GH2O values from 106 species of living birds and crocodilians were analyzed for this study, increasing the previous sample size of species known from morphometric data (Hoyt et al., 1979) by 85. Taxa selected were limited to those within the approximately 290 species for which experimental GH2O values were available in the literature, and, within these, to species for which morphometric data could be collected on their eggshells. A further limitation with selection of archosaur species for morphometric data was that species with pore canals that approximate simple (non-branching), straight structures were used because

14 the morphometric method using Fick’s first law assumes tubular pores for estimation of morphometric GH2O (Equation 2-2). Species with extremely complex pore canal structures [e.g., (Casuarius Brisson, 1760), (Dromaius Vieillot, 1816), (Pterocnemia G.R. Gray, 1871), Rhea (Rhea Brisson, 1760), and (Struthio L., 1758); Tyler and Simkiss, 1959; Board and Tullett, 1975; Board et al., 1977; Tyler and Fowler, 1979] were thus excluded from the dataset. Regardless of these limitations for species selection, this study includes a broad taxonomic range within Aves (>14 orders), with a large range of egg sizes [2.7 g for the smallest egg of the (Passer domesticus (L., 1758)) to 350.2 g for the largest egg of the Brown (Apteryx australis Shaw, 1813): Ar et al., 1974; Ar and Rahn, 1978, 1985; Calder et al., 1978] and with various nest types (e.g., covered and open nests) and nesting habitats, including extremes such as desert, montane, and polar regions.

2.2.2 Experimental water vapor conductance (Gex) −1 −1 Values (n = 106) of experimental GH2O (Gex: mg H2O∙day ∙Torr ) compiled from the literature were examined. All values had been calculated from experiments under known temperature (in most cases 25 °C or 298 °K) and relative humidity (0%) in desiccators (or sometimes in incubators with known relative humidity) with known barometric pressure (in most cases adjusted to 760 Torr) for at least several days. For this method, saturated water vapor pressure at the temperature of the desiccators and water vapor pressure surrounding the eggs in desiccators are equivalent to Pegg and Pnest, respectively (Ar et al., 1974). In most cases, water vapor pressure surrounding the eggs in desiccators is zero and thus Gex is equal to MH2O divided by saturated water vapor pressure at the temperature of the desiccators (in most cases 298 °K)

(Equation 2-1). Since more than one Gex value was usually available for each species, a mean value was calculated to represent conductance through the eggshell.

2.2.3 Morphometric water vapor conductance (Gmo) −1 −1 Values (n = 106) of morphometric GH2O (Gmo: mg H2O∙day ∙Torr ) were estimated using an equation (Equation 2-2) derived from Fick’s first law of gas diffusion (Ar et al., 1974). The right side of Equation 2-2 consists of three parts: assumed parameters of the nesting −1 environment (i.e., T and DH2O), measurements of the pore canals (i.e., Ap∙Ls , referred to as eggshell porosity in this study as in Zimmerman and Hipfner, 2007), and constants [i.e., c (=

15

9 −1 −1 7 3 −1 −1 −1 1.56·10 mg H2O∙s∙day ∙mol ) and R (= 6.24·10 mm ∙Torr∙mol ∙°K )]. Although Ap∙Ls differs among species, T and DH2O are generally assumed to be consistent among species (e.g., Ar et al., 1974; Ar and Rahn, 1985; Massaro and Davis, 2005; Jaeckle et al., 2012). This is a reasonable assumption because minor differences in T or DH2O do not strongly affect values of

Gmo. For example, previous papers used either 298 or 303 °K to calculate Gmo, which accounts for less than 15% of variation in Gmo (calculated here based on Equation 2-2 and values of T and

DH2O provided by Seymour, 1979 and Varricchio et al., 2013). 2 −1 To calculate Gmo for this study, 298 °K and 25.2 mm ∙s were used for the values of T and DH2O, respectively, as 298 °K was almost always used for estimation of Gmo (e.g., Ar et al.,

1974; Carey et al., 1990; Massaro and Davis, 2005; Jaeckle et al., 2012), as well as of Gex (e.g., Ar and Rahn, 1985). Using these values, Equation 2-2 can be simplified as

Ap [2-3] Gmo  2.1 Ls 2 Total pore area in an egg (Ap: mm ) is determined by multiplying mean individual pore 2 2 −2 area (A: mm ), eggshell surface area (As: mm ), and pore density (D: number of pores∙mm ) (Table 2.2; Figures 2.1a and 2.1b). Eggshell surface area was calculated from maximum egg length (L: mm) and breadth (B: mm), using equations of Paganelli et al. (1974) and Hoyt (1979)

(Table 2.2). Pore length (Ls: mm) was considered equal to eggshell thickness (Ar et al., 1974; Ar and Rahn, 1985). −1 For consistency, Ap∙Ls was recalculated here for all species (except for Red-winged Blackbird (Agelaius phoeniceus (L., 1766)) and Brown-headed (Molothrus ater

(Boddaert, 1783)) using A, B, D, L, and Ls from published data and from new measurements of eggshells. When D was not available, Ap was calculated by multiplying A and N, the latter of which is the total number of pores in the egg (Table 2.2). In most cases, values of each variable

(i.e., A, B, D, L, Ls, and N) were gathered from different sources for a single species as in

Zimmerman and Hipfner (2007). Because more than one value of A, B, D, L, Ls, and N was usually available for eggs of each species, a mean value of each variable was calculated. Values of Ls for Australian Brush- (Alectura lathami J.E. Gray, 1831) and (Leipoa ocellata Gould, 1840), taken from the literature, were averaged between unincubated and hatched eggshells because Ls of megapode species is known to significantly decrease through incubation due to absorption by embryos, thus increasing GH2O (Booth and Seymour,

16

1987; Booth and Thompson, 1991). A mean value is likely appropriate because it provides a better representation of thickness of megapode eggshells during incubation. For the Superb (Menura novaehollandiae Latham, 1802), A was calculated from the pore canal diameter provided by Lill (1987), assuming that a cross section of a pore canal is round. Values of A for Agelaius phoeniceus and Molothrus ater could not be obtained from previous studies, so −1 Ap∙Ls values were taken from Jaeckle et al. (2012).

Methods used for new measurements for A, D, and Ls were comparable with previous studies that estimated pore sizes and density of living or fossil archosaur eggs (e.g., Tullett and Board, 1977; Hoyt et al., 1979; Seymour, 1979; Ar and Rahn, 1985; Deeming, 2006). To estimate A, tangential thin sections of eggshells were photographed using a petrographic (Leica DM2500P) and cross-sectional pore area was analyzed with Adobe Photoshop version 12.0.1 and Image J (W. Rasband, National Institutions of Health, Bethesda, Maryland, USA). Pore density was estimated from the number of pore openings on the outer surface of the eggshell under a binocular (Leica M80), following the pore count procedure described by Ar and Rahn (1985). For samples of bird eggshell, the inner surface was stained with methylene blue solution to identify small pore openings on the outer surface. In order to minimize the regional variation of D within an egg, eggshell samples were taken from different parts of the egg (i.e., around equator and two poles) whenever possible. Eggshell thickness was measured with the digital micrometer Mitutoyo CPM30-25MJ (precision = 2 μm). Eggshell samples measured for this study were taken from the egg collections of the Hokkaido University Museum (Sapporo, Hokkaido, Japan), Royal Ontario Museum (Toronto, Ontario, Canada), Yale Peabody Museum (New Haven, Connecticut, USA), and Zelenitsky Egg Catalogue (Calgary, Alberta, Canada) (Appendix 2.1).

2.2.4 Egg mass

Egg mass (M: g) of the 106 species for which Gex and Gmo were estimated was compiled for this study. All values for egg mass were taken from previous studies, which weighed either fresh eggs or non-fresh raw eggs. Mass of non-fresh raw eggs was estimated in previous papers by refilling the air cell with water (an experimental estimation method developed by Grant et al., 1982a). When more than one value of M was available for eggs of each species, a mean value was calculated.

17

2.2.5 Statistical analyses For statistical analyses, the dataset of all species (n = 106) and those of separate taxonomic orders, including Anseriformes (n = 33), Charadriiformes (n = 25), and Galliformes (n = 13), were examined. Orders were analyzed to show if datasets of all species and individual orders provide comparable results. Other taxonomic orders in the dataset were not analyzed because of their small sample sizes (less than five species).

Values of Gex and Gmo for the datasets of all species and the three orders were compared using both phylogenetic and nonphylogenetic statistical analyses. The phylogenetic approach, which considers phylogenetic relationships of the species within the dataset, is used for bivariate analyses to avoid inflation of the rate of type I errors, which can be caused by the assumption of the independence of data points (i.e., species) (e.g., Garland et al., 2005). In this study, the phylogenetic approach was implemented for regression analyses and correlation tests, except for the Passing–Bablok linear regression that uses a non-phylogenetic approach, to estimate phylogenetically corrected regressions and correlation coefficients. A non-phylogenetic (i.e., conventional) approach was also implemented to compare results with those of the phylogenetic approach.

Prior to statistical analyses, values for experimental and morphometric GH2O, as well as egg mass, were transformed into base-10 logarithms (i.e., log Gex, log Gmo, and log M), because logarithmic distribution is commonly present in nature, normalizes distribution, and reduces heteroscedasticity and the effects of outliers (e.g., Zar, 1968; Sokal and Rohlf, 1995). Although most data points were averaged from multiple values with different sample sizes, it was assumed that data points were equally weighed. Conventional statistical analyses were conducted with MedCalc Statistical Software version 13.0 (MedCalc Software bvba, Ostend, Belgium; http://www.medcalc.org, accessed 3 June 2014). Phylogenetically corrected approaches, including reconstruction of a phylogenetic tree, were implemented using the PDAP module version 1.16 (Midford et al., 2010) of the software Mesquite version 2.75 (Maddison and Maddison, 2010).

2.2.5.1 Conventional statistical approach Descriptive statistics (i.e., mean, standard error, standard deviation, and range) were

18 examined for each variable (i.e., log Gex and log Gmo) of the datasets for all species and the three orders.

For each dataset, bivariate plots were created between log Gex and log Gmo, between log

Gex and log M, and between log Gmo and log M. Ordinary least-squares (OLS) regressions were used for the plots of log GH2O (both log Gmo and log Gex) against log M. Correlations between log

Gex and log Gmo were analyzed using Pearson’s correlation coefficients (r). Passing–Bablok linear regression analysis (Passing and Bablok, 1983) was used to detect systematic errors (i.e., constant error and proportional error) between log Gex and log Gmo. This analysis reveals a constant error when the 95% confidence interval of the intercept does not include zero and a proportional error when the confidence interval of the slope does not include one. A Bland–Altman plot (Bland and Altman, 1986) was used to graphically evaluate the agreement between the datasets of the experimental and morphometric methods. The agreement between the two methods is determined by plotting the differences of each pair of variables (log

Gex – log Gmo) against the mean values of two variables [0.5·(log Gex + log Gmo)] for each species. Because linear relationships on the plot may indicate proportional errors between the two methods (Bland and Altman, 1995), correlations between the differences and the mean values were tested. Also, the 95% limits of agreement, which are used to evaluate how much two methods agree or differ, were calculated. If the range of these limits is small enough, the two methods are interchangeable. Lower and upper 95% limits of agreement are calculated as [2-4] d 1.96  SD for the lower limit of agreement and [2-5] d 1.96  SD for the upper limit of agreement, where is the mean difference and SD is the standard deviation of the mean difference (Altman and Bland, 1983). Because this approach assumes normal distribution for the dependent variable (log Gex − log Gmo), Shapiro–Wilk test for non-normality was implemented.

2.2.5.2 Phylogenetically corrected statistical approach In addition to the conventional approach, a phylogenetically corrected approach was used for bivariate analyses, using the technique of independent contrast (Felsenstein, 1985). In

19 this approach, phylogenetically-generalized least-squares (PGLS), which assumes an evolutionary process with Brownian motion (e.g., Lavin et al., 2008), was implemented for testing regressions and correlations. The phylogenetically corrected approach was not applied for Passing–Bablok linear regression because it is used for a non-phylogenetic method only. For a phylogenetically corrected approach, a tree of 106 species of living birds and crocodilians was reconstructed using Hackett et al. (2008) for a large-scale phylogeny and additional papers for inter-relationships within each (Appendix 2.4). Because the tree topology was compiled from various sources, an arbitrary branch length model was used (Garland et al., 1992). Adequately standardized branch lengths of the tree were obtained when all branch lengths were set to one.

2.3 RESULTS 2.3.1 Descriptive statistics

Both experimental and morphometric GH2O values were compiled for the eggs of 106 living archosaur species (Appendices 2.2 and 2.3). Descriptive statistics for all species and for three orders (i.e., Anseriformes, Charadriiformes, and Galliformes) are provided in Table 2.3.

Datasets for all species and the three orders show that values of log Gmo have wider ranges than those of log Gex and that the mean values were higher for log Gex values than for log Gmo values (Table 2.3).

2.3.2 Bivariate plots of GH2O

Bivariate plots between log Gex and log Gmo, between log Gex and log M, and between log Gmo and log M indicate positive linear relationships (Figures 2.2a–2.2d and 2.3a–2.3d).

Pearson’s correlation coefficients revealed that log Gex and log Gmo are moderately to highly correlated in all datasets (r > 0.63 for conventional approaches and r > 0.66 for phylogenetically-corrected approaches). Also, both log Gmo and log Gex are correlated with log M in all datasets, although the slopes of the regressions are steeper in log Gmo than in log Gex, which are shown by both conventional and phylogenetically-corrected approaches (Table 2.4).

The bivariate plots that include all species indicate that log Gmo tends to be underestimated relative to log Gex when values of log Gex or log M are low (Figures 2.2a and

2.3a). In the bivariate plot of log Gex vs. log Gmo, 69% of the data points lie below the line of log

20

Gex = log Gmo, particularly species with lower log Gex values. Similarly, values of log Gmo are underestimated relative to those of log Gex at lower values of log M (e.g., 79% of species underestimated when log M < 1.5), whereas values of log Gmo are similar to those of log Gex at higher values of log M. In bivariate plots of the three orders (Figures 2.2b–2.2d and 2.3b–2.3d), species with values of low log Gex (< 0.75) and low log M (< 1.5) are limited (n ≤ 7) so the underestimation of log Gmo is not apparent.

2.3.3 Passing–Bablok linear regression A Passing–Bablok linear regression analysis revealed that there are both proportional and constant errors between log Gmo and log Gex for the dataset of all species (Table 2.4; Figure 2.4a). The 95% confidence interval of the slope does not include 1.0, suggesting that a proportional error exists between these two methods. Also, a constant error between the two methods was indicated as the 95% confidence interval of the intercept did not include zero. Both constant and proportional errors were also detected for the datasets of Anseriformes and Galliformes, but only a constant error was found in the dataset of Charadriiformes (Table 2.4; Figures 2.4b–2.4d).

2.3.4 Bland–Altman plot

The Bland–Altman plots show comparisons of the GH2O datasets obtained by experimental and morphometric methods (Table 2.5; Figures 2.5a–2.5d). Distributions of the differences of log Gex and log Gmo were assumed to be normal distributions in all datasets (p >> 0.05, except one dataset where p > 0.01). For the dataset of all species, the mean of the differences between log Gex and log Gmo was 0.15, indicating that Gex values are, on average, 1.4 times (i.e., antilog of 0.15) higher than Gmo values. The lower and upper 95% limits of agreement for the dataset of all species were -0.40 and 0.69, respectively. The antilog of these limits (0.40 and 4.92) indicates that for 95% of cases, the Gmo values will be between 0.40 and 4.92 times the values of Gex, suggesting that Gex does not accurately predict Gmo. Similar ranges of the lower and upper 95% limits of agreement were found in the datasets of the three orders (Table 2.5). Also, the Bland–Altman plot of the dataset of all species shows a negative relationship between the differences and the mean values of log Gex and log Gmo (Figure 2.5a) with both conventional and phylogenetically-corrected approaches (r = -0.60 for conventional approach and r = -0.52 for

21 phylogenetically-corrected approach: p < 0.01). Similar negative correlations were also found on the Bland–Altman plots of the three orders (Table 2.5; Figures 2.5b–2.5d).

2.4 DISCUSSION A large increase in the morphometric dataset in this study allowed rigorous statistical comparisons between the datasets of the morphometric and experimental GH2O methods of living archosaur species. Both conventional and phylogenetically corrected approaches indicate that log

Gex is significantly correlated to log Gmo, although there is disagreement between the datasets of these two methods likely due to constant and/or proportional errors. These results were consistent regardless of the dataset (i.e., all species, Anseriformes, Charadriiformes, and Galliformes), except a proportional error was not found in the Charadriiformes dataset.

Disagreement between the datasets of the two methods is most apparent when log Gex or log M values are low (Figures 2.2a–2.2d and 2.3a–2.3d). The species with lower log Gex (< 0.75) or lower log M values (<1.5) tend to have underestimated log Gmo values compared with species with higher log Gex or higher log M values (Figures 2.2a–2.2d and 2.3a–2.3d). This underestimation is more apparent for the dataset of all species than for the datasets of the three orders (i.e., Anseriformes, Charadriiformes, and Galliformes), perhaps because the number of species with lower log Gex or log M are limited in the datasets of these orders. Although difficult to ascertain, the reason(s) for the disagreement between the two methods could be due to a variety of errors potentially related to the collection or treatment of the Gex and Gmo data. One reason for the disagreement could be related to sampling or measurement errors primarily among eggs with lower mass or lower Gex. With respect to the experimental method, estimations are based solely on repeat measurements of egg mass under known conditions (i.e., temperature, humidity, and barometric pressure), so there should be few sources of systematic error that would cause such a discrepancy. However, one possible methodological error could be associated with cracks developing in the eggshell of smaller eggs. Smaller eggs are extremely fragile because of thin eggshell (Ar et al., 1979). Thus, cracks could have developed accidentally during sampling or experimentation, which would allow extra water loss from an egg. This would particularly affect species with small eggs and result in an erroneously high Gex. Unfortunately, it is not possible to verify such a methodological error because all Gex values were gleaned from the literature.

22

With respect to the morphometric method, Gmo values could be affected by measurement errors of several eggshell features that are used for the Gmo estimation (Equation 2-2), including single pore area, pore density, surface area of the egg, and eggshell thickness

(equal to pore length). The lower Gmo values in small eggs could result from underestimation of single pore area or pore density, or overestimation of pore length, although it is difficult to explain such an occurrence in mainly small eggs because methods used for measuring these parameters are similar among studies regardless of egg size. Other possible errors related to measurements in small-bird eggs could be that assumptions for estimating eggshell porosity (Ar et al., 1974; Paganelli et al., 1974; Hoyt, 1979; Ar and Rahn, 1985) are potentially erroneous, including (i) pore length is equal to eggshell thickness, (ii) pore canals are simple cylinders, and (iii) eggshell surface area is estimated with equations using egg length and breadth. Although it has not been investigated if the first assumption is relevant for various bird species, the latter two have been shown to be reasonable in bird eggs with lower mass and Gex because shorter pores (≈ smaller eggs) tend to be more cylindrical (Toien et al., 1988) and the equations of surface area are applicable regardless of egg size or egg shape in birds (Hoyt, 1979; Paganelli et al., 1974). Another reason for disagreement between the two methods could be the use of various sources and samples to compile the Gex and Gmo data. For most species, Gex estimations were based on eggs different from those for Gmo estimations because measurements were obtained from different papers or samples; thus, mean values for Gex and Gmo for each species were calculated from separate samples. If, however, use of different samples was the source of error, the tendency for large discrepancies in eggs with lower log M or log Gex is difficult to explain because errors would be expected to occur randomly regardless of egg size or Gex value. Although various possibilities for the disagreement between the dataset of the two methods are discussed, there is no good explanation for the discrepancy or the source of errors. Regardless, the discrepancy revealed here should be a caution when using the morphometric method to estimate GH2O, as Gmo may not be a reliable proxy for experimental values (on average,

40% difference), particularly for eggs with lower mass and lower Gex. Further research is required to investigate possible sources of errors in the experimental and morphometric methods leading to the discrepancy.

Because experimental GH2O cannot be measured for fossil eggs, the morphometric method has been used to estimate GH2O for dinosaurs (Seymour, 1979; Williams et al., 1984;

23

Sabath, 1991; Mou, 1992; Grigorescu et al., 1994; Mikhailov et al., 1994; Sahni et al., 1994; Antunes et al., 1998; Deeming, 2006; Jackson et al., 2008; Grellet-Tinner et al., 2012; Varricchio et al., 2013; Zhao et al., 2013). As a result, direct comparison between Gmo values of dinosaurs and Gex values of living archosaurs has been used to predict nest type in dinosaurs because Gex values in living species appear to correspond to nest type (Seymour, 1979). As revealed here, however, Gmo does not necessarily reflect Gex, so caution should be used if comparing Gmo values of dinosaurs with Gex values of living archosaurs.

Based on discrepancies between Gex and Gmo, it is reasonable that an alternative approach be used for predicting nest type in extinct species, one that does not rely on comparisons between Gex and Gmo. For the Gmo equation (Equations 2-2 and 2-3), T and DH2O are assumed to be consistent among species (e.g., Ar et al., 1974; Ar and Rahn, 1985; Massaro and Davis, 2005; Jaeckle et al., 2012) because differences of these variables are assumed to be −1 negligible (see also the Materials and Methods in this chapter), so eggshell porosity (Ap∙Ls ) is the only variable that varies among species (Equation 2-3). Since eggshell porosity directly affects the water vapor diffusion of eggs (Ar et al., 1974), this variable could be a suitable alternative to Gmo and potentially be a straightforward predictor of nest type. Previous studies were not able to directly compare Gmo or eggshell porosity between living and fossil taxa because eggshell porosity data were limited for living species. Here eggshell porosity was estimated for a large number of living archosaur species (n = 106), which will allow for statistical comparisons of eggshell porosity between fossil and living species. If relative values of eggshell porosity are shown to differ among nest types (i.e., covered and open nests) in living archosaur species, then this variable may be a suitable alternative to GH2O to predict nest type in dinosaurs.

24

Table 2.1. Comparisons of two methods for estimating water vapor conductance of archosaur eggs. *See text for abbreviations. Method Experimental method Morphometric method Equation* M H 2O c DH 2O Ap [2-1] GH 2O  [2-2] GH 2O   Pegg  Pnest R T Ls

Abbreviation Gex Gmo Estimated by Weighing egg mass daily under known Calculating eggshell pore geometry as temperature and humidity in the well as assumed incubation conditions laboratory and constants Applied for Only living birds and crocodilians Most often extinct taxa, such as dinosaurs and fossil birds

25

Table 2.2. Lists of variables (A) and equations (B) used for estimation of eggshell porosity (variables and definitions from Deeming, 2006 and references therein). (A) Variable Definition Unit A Mean individual pore area (μm2) 2 Ap Total pore area (mm ) −1 Ap∙Ls Eggshell porosity (mm) 2 As Surface area of eggshell (mm ) B Maximum egg breadth (mm) D Pore density (mm−2) L Maximum egg length (mm)

Ls Shell thickness (i.e., pore (mm) length) M Egg mass (g) N Total number of pores (per egg) V Egg volume (mm3)

(B) Equation Sources

Ap = A · As · D = A · N Seymour (1979) 0.666 As = 4.951 · V Paganelli et al. (1974) M = 5.48 · 10−4 · L · B2 Hoyt (1979) V = 0.51 · L · B2 Hoyt (1979)

26

Table 2.3. Results of descriptive statistics for GH2O datasets.

Note that the mean log GH2O is higher in the datasets of the experimental method than the morphometric method. Abbreviations: CI, confidence interval; n, sample size; SD, standard deviation; SE, standard error. Variable Sample N Mean SE 95% CI SD Range log Gex All species 106 0.983 0.036 0.911 to 0.375 -0.084 to 1.055 1.991 Anseriformes 33 1.162 0.039 1.082 to 0.226 0.663 to 1.242 1.531 Charadriiformes 25 0.937 0.055 0.825 to 0.273 0.280 to 1.050 1.365 Galliformes 13 0.978 0.117 0.723 to 0.423 0.072 to 1.234 1.674 log Gmo All species 106 0.838 0.052 0.735 to 0.538 -1.051 to 0.942 1.993 Anseriformes 33 1.026 0.055 0.914 to 0.316 0.063 to 1.138 1.586 Charadriiformes 25 0.775 0.074 0.623 to 0.369 -0.060 to 0.928 1.298 Galliformes 13 0.915 0.171 0.543 to 0.616 -0.330 to 1.287 1.993

27

Table 2.4. Results of regression analyses by the conventional model, phylogenetically corrected model, and Passing–Bablok method. Note that * and † indicate proportional (*) and constant errors (†) exist between the datasets of the experimental and morphometric methods. Abbreviations: CI, confidence interval; n, sample size; OLS, ordinary least-squares; PB, Passing–Bablok regression; PGLS, phylogenetically-generalized least-squares; r2, coefficient of determination. Dependent Independent Model Sample n Slope CI of slope Intercept CI of intercept r2 variable variable log Gex log M OLS All species 106 0.857 0.776 to 0.938 -0.466 -0.607 to -0.325 0.809 Anseriformes 33 0.879 0.681 to 1.077 -0.499 -0.876 to 0.122 0.725 Charadriiformes 25 0.850 0.716 to 0.984 -0.465 -0.689 to 0.241 0.882 Galliformes 13 1.052 0.787 to 1.317 -0.796 -1.253 to 0.338 0.874 PGLS All species 106 0.820 0.720 to 0.920 -0.175 -0.442 to 0.093 0.717 Anseriformes 33 0.813 0.551 to 1.074 -0.282 -0.780 to 0.216 0.566 Charadriiformes 25 0.905 0.768 to 1.042 -0.635 -0.892 to -0.378 0.894 Galliformes 13 0.896 0.313 to 1.478 -0.457 -1.642 to 0.728 0.623 log Gmo log M OLS All species 106 1.130 0.980 to 1.279 -1.073 -1.332 to -0.813 0.683 Anseriformes 33 0.897 0.483 to 1.311 -0.668 -1.455 to 0.119 0.386 Charadriiformes 25 0.997 0.693 to 1.302 -0.870 -1.380 to 0.359 0.666 Galliformes 13 1.339 0.712 to 1.965 -1.343 -2.424 to 0.262 0.668 PGLS All species 106 0.955 0.754 to 1.156 -0.418 -0.954 to 0.118 0.462 Anseriformes 33 1.148 0.477 to 1.818 -1.124 -2.401 to 0.153 0.283 Charadriiformes 25 0.984 0.598 to 1.371 -0.701 -1.427 to 0.025 0.556

28

Dependent Independent Model Sample n Slope CI of slope Intercept CI of intercept r2 variable variable Galliformes 13 1.277 0.043 to 2.511 -1.234 -3.744 to 1.275 0.429 log Gmo log Gex PB All species 106 1.454 1.301 to 1.641* -0.587 0.802 to -0.427† Not measured Anseriformes 33 1.534 1.009 to 2.310* -0.820 -1.733 to -0.138† Not measured Charadriiformes 25 1.365 0.990 to 1.860 -0.504 -1.055 to -0.151† Not measured Galliformes 13 1.493 1.090 to 2.525* -0.579 -1.529 to -0.239† Not measured

29

Table 2.5. Results of the Bland–Altman plots between differences of the experimental (log

Gex) and morphometric (log Gmo) methods and their mean values.

Note that there are negative correlations between the differences and averages (* for p < 0.01 and

** for p < 0.05). Abbreviation: r, correlation coefficient.

Samples Mean Lower and upper 95% r by a r by a

difference limits of agreement conventional phylogenetic

approach approach

All species 0.145 -0.402 and 0.692 -0.603* -0.519*

Anseriformes 0.136 -0.345 and 0.617 -0.404** -0.709*

Charadriiformes 0.162 -0.283 and 0.607 -0.445** -0.433**

Galliformes 0.064 -0.535 and 0.662 -0.649** -0.691*

30

Figure 2.1. Porosity of archosaur eggshell.

(a) Generalized archosaur eggshell with pore canals, indicating individual pore area (A), eggshell surface area (As), maximum egg breadth (B), pore density (D), maximum egg length (L), and pore length (Ls). (b) Tangential thin section of eggshell of the (Pavo cristatus L.,

1758) with pores indicated by arrows.

31

Figure 2.2. Bivariate plots of experimental (log Gex) and morphometric (log Gmo) conductance values showing positive correlations between these variables: (a) all species, (b)

Anseriformes, (c) Charadriiformes, and (d) Galliformes.

2 The solid lines (y = x) indicate ideal regressions (r = 1.0) when values of log Gex and log Gmo are identical. Note that the grey highlighted region (log Gex < 0.75) of panel a indicates where log Gmo of most species (78% of all species) tends to be underestimated, whereas the underestimation is not apparent on the grey highlighted regions of panels b, c, and d.

32

Figure 2.3. Bivariate plots of water vapor conductance (log GH2O) and egg mass (log M) showing positive correlations between these variables: (a) all species, (b) Anseriformes, (c)

Charadriiformes, and (d) Galliformes.

The grey circles and solid lines indicate data points and the ordinary least-squares (OLS) regressions of experimental conductance values (log Gex), respectively, whereas the white circles and broken lines indicate those of morphometric conductance values (log Gmo). Note that the grey highlighted region (log M < 1.5) of panel a indicates where log Gmo of most species (79% of all species) are underestimated, whereas the underestimation is not apparent on the grey highlighted regions of panels b, c, and d.

33

Figure 2.4. Passing–Bablok regression plots for the datasets of experimental (log Gex) and morphometric (log Gmo) conductance values in living archosaurs: (a) all species, (b)

Anseriformes, (c) Charadriiformes, and (d) Galliformes.

Note that Passing–Bablok regressions (solid lines) with the 95% confidence intervals (broken lines) indicate evidence of proportional and/or constant errors. The grey lines indicate complete agreement between log Gex and log Gmo (y = x).

34

Figure 2.5. Bland–Altman plot of differences and mean values between experimental (log

Gex) and morphometric (log Gmo) conductance values: (a) all species, (b) Anseriformes, (c)

Charadriiformes, and (d) Galliformes.

Note that there are negative correlations between the differences and the mean values. Also note that wide ranges of upper and lower limits of agreement (broken lines) indicate weak agreement between these two methods. The solid lines indicate the mean values of the differences.

35

CHAPTER 3: EGGSHELL POROSITY PROVIDES INSIGHT ON EVOLUTION

OF NESTING IN DINOSAURS

3.1 INTRODUCTION

Nests are varied structures that play an important role in archosaur biology because they are used for incubating eggs and, in many species, for raising young. The nests can consist of simple scrapes or holes in the ground, bowl-shaped structures, or large vegetation mounds

(Hansell, 2000; Brazaitis and Watanabe, 2011), and their architecture is suited not only for the incubation of eggs in a given environment but also for the incubation behavior/method of a species.

Among extant archosaurs, two general types of nest are observed: 1) covered nests, in which the eggs are covered by organic/inorganic matter, are built by species that incubate their eggs using external heat sources (e.g., solar heat, plant decomposition, or geothermal heat: Deeming, 2002a), and 2) open nests, in which the eggs are not covered by substrate and left exposed, are built by species that brood their eggs. Because all crocodilian species build covered nests and all bird species, except those of megapodes, incubate eggs in open nests (Coombs, 1989), the transition from covered to open nest type likely occurred among non-avian dinosaurs (e.g., Varricchio et al.,

1997).

Nest types and associated nesting behaviors are poorly understood in extinct archosaurs, including non-avian dinosaurs, in part because nest structures and nesting materials are rarely preserved (Carpenter, 1999; Chiappe et al., 2004). Even on the rare occasions where nest structures are found (e.g., excavations, mounds; Varricchio et al., 1999; Chiappe et al., 2004;

Zelenitsky and Therrien, 2008b; Vila et al., 2010b), there is no indication of whether the eggs were covered by organic/inorganic material or surrounded by nesting materials as typically found in 36

living archosaurs. Consequently, other evidence related to egg clutches, such as their taphonomic and sedimentologic setting or eggshell structures (i.e., pore canals), have been used to infer the nest type of dinosaurs (e.g., Seymour, 1979; Chiappe et al., 2004).

Most prior studies have used a method that estimates the diffusive capacity of the eggshell, referred to as water vapor conductance (GH2O), to infer the types of nests built by dinosaurs (e.g.,

Seymour, 1979; Williams et al., 1984; Deeming, 2006; Jackson et al., 2008; Varricchio et al.,

2013). Water vapor conductance in living archosaurs has usually been measured experimentally via daily water loss of a fresh egg (e.g., Ar et al., 1974; Ar and Rahn, 1985). A theoretical formula to calculate water vapor conductance from eggshell porosity was also developed based on Fick's law of diffusion (herein referred to as morphometric GH2O) (Ar et al., 1974). This formula was used initially by Seymour (1979) to calculate GH2O for dinosaur eggs, and nest type was inferred on the premise that covered nests are found in living species with high GH2O and open nests in species with low GH2O values. Thus, morphometric GH2O values of dinosaurs were compared directly with experimental GH2O values of living archosaurs (e.g., Seymour, 1979; Williams et al., 1984;

Deeming, 2006; Jackson et al., 2008; Varricchio et al., 2013), although the latter were calculated from measurement of daily water loss of an egg (experimental GH2O) and not from the theoretical formula (morphometric GH2O). However, a recent study compared morphometric and experimental GH2O values in living archosaur species, and demonstrated that these two datasets/methods are mutually incongruent, likely due to systematic errors (Tanaka and Zelenitsky,

2014a). Thus direct comparison between morphometric and experimental GH2O values, as widely applied to infer nest type of dinosaurs, may not be valid.

As a viable alternative to the water vapor conductance method, I present a statistically rigorous approach using eggshell porosity in order to predict nest type in extinct archosaurs. This approach is applied to the eggs of a variety of dinosaurs, including titanosaurs, the theropod 37

Lourinhanosaurus, oviraptorosaurs, and troodontids, in order to assess their nesting habits and discuss the evolution of nest type and incubation behaviors among archosaurs.

3.2 MATERIALS AND METHODS

A series of methodological steps were taken to document the relationship between eggshell porosity and nest types in crocodylians and birds in order to infer nest types in extinct

−1 archosaurs. Data on eggshell porosity (Ap∙Ls , in mm), egg mass (M, in g), and nest types (see

"3.2.3 Nest classification for extant taxa") were compiled for living crocodylians and birds and subsequently compared statistically to test whether eggshell porosity relative to egg mass differs between open and covered nest types. Eggshell porosity and egg mass were then estimated for a variety of extinct archosaurs, including crocodylomorphs, non-avian dinosaurs, and birds.

Through comparison with the extant dataset, discriminant analyses were used to infer nest types in extinct taxa. Both phylogenetic and non-phylogenetic (i.e., conventional) approaches were applied for the statistical analyses.

3.2.1 Relationship between water vapor conductance and eggshell porosity

Water vapor conductance of living species has usually been measured experimentally

(e.g., Ar et al., 1974; Ar and Rahn, 1985), but it has also been shown to be related to the geometry of eggshell pore canals. Ar et al. (1974) were the first to derive a mathematical equation to calculate morphometric water vapor conductance (GH2O) using pore geometry in archosaur eggs.

This equation is expressed as:

c DH 2O Ap [3-1] GH 2O   R T Ls 9 −1 −1 where c is a unit conversion constant (= 1.56·10 mg H2O∙s∙day ∙mol ), DH2O is the diffusion

38

coefficient of water vapor (mm2∙s−1) in air, R is the universal gas constant (= 6.24·107

3 −1 −1 −1 mm ∙Torr∙mol ∙°K ), T is the absolute temperature of incubation (°K), Ap∙Ls is eggshell

2 porosity, Ap is the total pore area of an egg (mm ), and Ls is pore length (mm) (Ar et al., 1974).

Since many variables (i.e., c, DH2O, R, and T) can be safely assumed to be consistent among species

(e.g., Ar et al., 1974; Ar and Rahn, 1985; Tanaka and Zelenitsky, 2014a), the equation can be

−1 simplified and expressed as GH2O = 2.1∙ Ap∙Ls (see Tanaka and Zelenitsky, 2014a). Morphometric water vapor conductance is thus directly proportional to eggshell porosity. Given that morphometric water vapor conductance (and hence eggshell porosity) is influenced by absolute nest humidity (Appendix 3.1), which in turn is correlated with nest architecture or type (covered vs. open, see Tanaka and Zelenitsky, 2014b), a correlation between eggshell porosity and nest types can be sought (see Appendix 3.1 for further explanation).

3.2.2 Selection of extant taxa

Eggshell porosity, egg mass and nest type for 127 species of extant birds and crocodylians were gathered from either the literature (see Tanaka and Zelenitsky, 2014a) or via new measurements of egg specimens (Appendices 3.3–3.5). Egg specimens were permitted to be accessed from the institutions listed in Appendix 3.3. The dataset includes only species with pore canals that approximate simple (unbranched) or tubular structures because porosity of eggshells with more complex pores (e.g., branched pores) could not be accurately estimated (e.g., Casuarius,

Dromaius, Pterocnemia, Rhea, and Struthio; Tyler and Simkiss, 1959; Board and Tullett, 1975;

Board et al., 1977; Tyler and Fowler, 1979). Although some pores of crocodylian eggshells may be irregularly shaped, they are usually simple and straight (Marzola et al., 2015) and are here assumed to be tubular.

39

3.2.3 Nest classification for extant taxa

Nest structures of extant archosaurs were classified into two general types, covered nests and open nests, based on information available in the literature (Appendix 3.5). Covered nests are defined as those in which the eggs are completely covered with vegetation and/or sediment (e.g., mound or infilled hole nests on/in the ground), whereas open nests are those in which the eggs are partly or fully exposed, and may have nest materials surrounding a portion of the eggs (e.g., scrape, cup, plate, and dome nests) (see Tanaka and Zelenitsky, 2014b).

Certain aquatic birds (e.g., Podicipediformes, Cygnus, Oxyura, Chlidonias niger, and

Gavia immer) were excluded from this study due to their unusual nesting style. Because these birds build open nests floating on water with nest materials that can be wet (Carey, 1980, 2002;

Davis et al., 1984; Dunn and Agro, 1995; McIntyre and Barr, 1997), presumably resulting in high nest humidity (Davis et al., 1984; Dunn and Agro, 1995; Carey, 2002; Tanaka and Zelenitsky,

2014b), their eggshell porosity and water vapor conductance are anomalously high for birds with open nests (Davis et al., 1984; Ar and Rahn, 1985; Davis and Ackerman, 1985).

3.2.4 Selection of fossil eggs/ootaxa

Eggshell porosity and egg mass for 29 extinct archosaur taxa and ootaxa (i.e., egg taxa) were compiled from the literature or from new/additional measurements of egg specimens listed in

Appendix 3.3. Only species and oospecies with simple pore canals and for which data for individual pore area, pore density, pore length, and egg length and breadth were available were included in this study (Table 3.1). Eggs and ootaxa with complex or irregular pores [e.g.,

Dendroolithidae (Torvosaurus and possibly therizinosaur), Faveoloolithidae (?Sauropodomorpha),

Ovaloolithidae (?), and Spheroolithidae (Maiasaura-like eggs, presumably hadrosaur)] and taxa/ootaxa for which the data were potentially derived from the combination of 40

multiple oospecies (e.g., '' and '' in Seymour, 1979 and

Elongatoolithidae in Sabath, 1991 and Mikhailov et al., 1994) were not included in this study. Also, eggshell porosity for enantiornithine eggs (e.g., 'Gobipteryx minuta' in Sabath, 1991) was not estimated in this study because the original article (Sabath, 1991) indicated questionable values for both total number of pores and individual pore area.

The taxonomic affinity of most ootaxa considered in this study is well-established, particularly at higher taxonomic levels. For example, the ootaxon Bauruoolithus is attributed to a crocodylomorph based on eggshell microstructure (Oliveira et al., 2011). The ootaxon

Megaloolithus patagonicus is referred to a titanosaur based on its association with embryonic remains (Chiappe et al., 1998, 2001), thus eggs of the Megaloolithidae oofamily are widely regarded as belonging to sauropods (Chiappe et al., 1998; Sander et al., 2008; Wilson et al., 2010).

Formerly classified in Megaloolithidae, the ootaxon Cairanoolithus has recently been re-assigned to a new oofamily, Cairanoolithidae, by Sellés and Galobart (2016) who suggested it may belong to an ornithischian dinosaur.

The taxonomic identity of some theropod eggs is also known based on association with either embryonic or parental skeletal remains. These include the eggs of Lourinhanosaurus antunesi, a large theropod of either allosauroid (Benson, 2010; Benson et al., 2010) or coelurosaurian (Carrano et al., 2012) affinity, the ootaxon yaotunensis, assigned to an oviraptorosaur (Cheng et al., 2008), and the ootaxon Prismatoolithus levis, assigned to Troodon formosus (Varricchio et al., 2002). Elongatoolithid and prismatoolithid ootaxa are attributed to

Oviraptorosauria (Dong and Currie, 1996; Clark et al., 1999; Norell et al., 2001; Sato et al., 2005;

Cheng et al., 2008; Weishampel et al., 2008; Fanti et al., 2012) and non-oviraptorosaur maniraptorans (Zelenitsky and Therrien, 2008a; Lopez-Martinez and Vincens, 2012), respectively, based on eggshell microstructure similarities with eggs of known taxonomic identity. The ootaxon 41

Continuoolithus is assigned to an indeterminate theropod based on egg and eggshell morphology

(Zelenitsky et al., 1996; Zelenitsky and Therrien, 2008a; Jackson et al., 2015). eggshells have been assigned to two small-bodied ( body masses 20–30 kg) species, Pachyornis geranoides and Euryapteryx sp., based on DNA analyses (Huynen et al., 2010).

3.2.5 Eggshell porosity

−1 Eggshell porosity (Ap∙Ls , in mm) of both living and extinct archosaurs was determined

2 by dividing the total pore area of an egg (Ap, in mm ) by pore length (Ls, in mm) (Tables 3.2 and

3.3; Figures 3.1a and 3.1b), except for Macroolithus yaotunensis by Wiemann et al. (2015) where eggshell porosity was calculated based on its morphometric GH2O value. Total pore area of an egg was calculated by multiplying individual pore area (A, in mm2) by pore density (D, in mm−2) and

2 eggshell surface area (As, in mm ). When pore density was not available in the literature, total pore area was calculated by multiplying individual pore area by the total number of pores in an egg (N).

Because more than one value was usually available for each variable (i.e., A, D, Ls, N, egg length, egg breadth), a mean value was calculated from the various sources/samples for each taxon/ ootaxon.

For taxa/ootaxa where eggshell porosity values could not be obtained from the literature, they were calculated from the measurement of relevant variables in thin sections of eggshell specimens. Whenever possible, eggshell samples were taken from different regions of an egg (i.e., around equator and two poles) in order to capture regional variation; the eggshell is slightly thicker and shows lower pore density at the pointed pole than at the blunt pole (e.g., Handrich, 1989;

Balkan et al., 2006). Shell thickness, which can be used as a proxy for pore length (Ar et al., 1974;

Ar and Rahn, 1985), was measured with a digital micrometer Mitutoyo CPM30-25MJ (precision =

2 μm). Individual pore area was measured from tangential thin sections of eggshells using a Leica 42

DM2500P petrographic microscope (Figures 3.1c–3.1e) following the procedures of Tanaka and

Zelenitsky (2014a). For pore counting in living birds, the inner surface of the eggshells was stained with methylene blue solution to accentuate the pores (see Ar and Rahn, 1985). Pore density was estimated by counting the number of pore openings on the outer surface of eggshells using a Leica

M80 binocular microscope, following the procedure of Tanaka and Zelenitsky (2014a). Pore density in Troodon was estimated from tangential thin sections because the outer surface was poorly preserved. Eggshell surface area was calculated from maximum egg length (L, in mm) and breadth (B, in mm), both obtained from the literature, using the equations of Paganelli et al. (1974) and Hoyt (1979) (Tables 3.2 and 3.3), except for the ootaxa andrewsi,

Macroolithus rugustus, and M. yaotunensis for which eggshell surface area were obtained from

Mou (1992). Because egg length and breadth of the Pachyornis geranoides and Euryapteryx sp. specimens studied are unknown, these values were taken from intact eggs (AIM LB4003, LB4005, and an unregistered AIM egg of Gill, 2006) found at the same locality, which show comparable eggshell thickness and pore morphology to the specimens used in this study (Archey, 1931, 1941).

A possible caveat for the calculation of porosity in fossil eggshells is that can alter pore dimensions (Jackson et al., 2008; Clayburn et al., 2004). Diagenetic dissolution for example can decrease pore length and enlarge pore canals, resulting in overestimation of eggshell porosity (Jackson et al., 2008; Clayburn et al., 2004). Because most values for fossil specimens were obtained from the literature it is impossible to assess the impact of diagenetic alteration on the ootaxa considered in this study. I proceed with the assumption that, overall, diagenesis did not significantly affect calculation of eggshell porosity, because there was no mention of diagenetic alternation in most papers.

43

3.2.6 Egg mass

Mean egg mass (M, in g) for living and extinct archosaurs was compiled for this study.

Egg mass for living species was obtained from the literature (Appendix 3.4) and that for fossil taxa/ootaxa was estimated from egg length and breadth using the equation of Hoyt (1979) (Table

3.3). Although other methods exist to estimate fossil egg mass (see Varricchio et al., 2013), they produce results that are consistent (within 10%) with Hoyt’s (1979) method (Varricchio et al.,

2013; Jackson et al., 2015). Therefore, for consistency, Hoyt's method was applied to all extinct ootaxa/taxa. Egg mass for Elongatoolithus andrewsi, Macroolithus rugustus, and M. yaotunensis was taken from Mou (1992), who had used Hoyt's (1979) method to derive his estimates.

3.2.7 Phylogenetic distribution of nest type

The nature of the phylogenetic distribution (i.e., random vs. clumped) of nest types among living archosaurs was investigated based on the compiled extant dataset in this study.

Because nest type can be coded as a binary trait (covered vs. open), Fritz and Purvis' (2010) D statistic was calculated by running 1000 permutations of the 'phylo.d' function of the package

'caper' using the software platform R3.1.3 (http://www.r-project.org/). For the D statistic, a value equal to or higher than 1.0 indicates a random phylogenetic distribution, whereas a value equal to or lower than 0 indicates a non-random phylogenetic distribution (i.e., phylogenetically clumped).

The 'phylo.d' function provides p values to indicate whether the estimated D statistic is significantly different from 0 and 1, respectively.

The D statistic was run using a phylogenetic tree of 127 species of living birds and crocodylians compiled from the large-scale phylogeny of Jarvis et al. (2014) and other publications for small-scale interrelationships (Appendix 3.8). Branch length was estimated from the divergence times of each node following the procedures of Motani and Schmitz (2011) and 44

Schmitz and Motani (2011). Divergence times of major clades were obtained from Time Tree

(http://timetree.org) for birds and from Oaks (2011) for crocodylians. Terminal taxon ages were set to zero. The phylogenetic tree and character matrix were constructed with the PDAP module v.1.16 (Midford et al., 2010) of the software Mesquite 3.02 (Maddison and Maddison, 2010).

3.2.8 Analysis of covariance

Eggshell porosity relative to egg mass was compared between extant open-nesting and covered-nesting archosaurs using both conventional and phylogenetically-corrected analysis of covariance (ANCOVA and pcANCOVA, respectively). Non-phylogenetic, ordinary least-squares

(OLS) regression was implemented for conventional ANCOVA with IBM SPSS Statistics v.

22.0.0 (IBM SPSS Inc.), whereas phylogenetically-corrected ANCOVA was implemented with the MATLAB (MathWorks Inc.) program Regressionv2.m (available upon request from T.

Garland Jr.) following the method of Lavin et al. (2008). A phylogenetic variance-covariance matrix for Regressionv2.m was generated with the DOS PDDIST program (Garland and Ives,

2000). Regressions for pcANCOVA were generated with two evolutionary models: regressions with Brownian motion (PGLS) and Ornstein-Uhlenbeck models (RegOU). PGLS assumes an evolutionary process with "random walk in continuous time" (e.g., Lavin et al., 2008), whereas

RegOU assumes an evolutionary process of "wandering back and forth on a selective peak"

(Felsenstein, 1988; Butler and King, 2004; Lavin et al., 2008). These three regression models

(OLS, PGLS, and RegOU) were compared using the Akaike Information Criterion (AIC) to determine the best fit model of regression, where a lower AIC value indicates a better fit (e.g.,

Lavin et al., 2008; Swanson and Garland, 2009; Gartner et al., 2010).

Nest type (open and covered nests) was considered a categorical variable, a covariate of egg mass, and a dependent variable of eggshell porosity in these analyses. Values of eggshell 45

−1 porosity (Ap∙Ls ) and egg mass (M) were log-10 transformed prior to analysis. The normality and homogeneity of variances of the dataset were tested by non-phylogenetic Shapiro-Wilk tests and

−1 Levene tests using IBM SPSS Statistics v. 22.0.0. Residuals of log Ap∙Ls , calculated from OLS regressions for each nest type, were used for the Shapiro-Wilk tests.

The phylogenetic tree compiled for the D statistic (see above) was used for the pcANCOVA. In addition to the branch length determination method based on divergence time used for the D statistic, an arbitrary standardized method was also applied to assign branch length for the pcANCOVA because branches were not adequately standardized by divergence time. An arbitrary branch length model was used by following the procedure of Garland et al. (1992), resulting in all branch lengths equal to one.

3.2.9 Discriminant analysis

Nest type of fossil taxa/ootaxa was inferred by analyzing their eggshell porosity using conventional, non-phylogenetic linear discriminant analysis (LDA) and the phylogenetic flexible discriminant analysis (pFDA) of Schmitz and Motani (2011). While LDA was applied to all extinct taxa and ootaxa examined, pFDA could only be used for fossil eggs of known taxonomic affinities (i.e., titanosaurs, Lourinhanosaurus, oviraptorosaurs, Troodon, and ) as knowledge of the precise phylogenetic relationships between taxa is required for this method. Linear discriminant analysis was implemented with IBM SPSS Statistics v. 22.0.0, whereas pFDA was conducted in R3.1.3 with the phylo.fda.v0.2.R script provided by L. Schmitz

(https://github.com/lschmitz/phylo.fda). LDA and pFDA were used to compare log-transformed values of eggshell porosity and egg mass of extinct archosaurs to those of living archosaurs

(grouped a priori into open and covered nests categories) to infer the nest type for each extinct taxon/ootaxon. In order to test if a phylogenetic bias affects the form-function relationship in the 46

dataset, the pFDA method provides an estimate of Pagel's lambda, where a lambda value of zero reveals no phylogenetic bias and a value of one indicates a strong bias where character evolution follows the Brownian motion model (Motani and Schmitz, 2011). The pFDA method also provides a series of predictions for each taxon as a function of changing lambda value from 0 to 1. Prior probabilities of nest types, which are required for discriminant analyses, were based on the proportions of open and covered nest types found in the dataset of living archosaurs since the proportion of each nest type in extinct archosaurs is unknown. The misclassification rate was calculated for both LDA and pFDA based on the proportion of erroneously classified species.

Since the misclassification rate of pFDA varies as a function of lambda values, the change in the overall misclassification rates through lambda values from 0 to 1 was also determined.

For pFDA, six extinct taxa were included in the composite phylogenetic tree of living archosaurs (Appendix 3.9). Because pFDA requires divergence times for estimation of branch length, phylogenetic relationships and divergence times of the extinct archosaur taxa were obtained from Bunce et al. (2009), Choiniere et al. (2010), Nesbitt (2011), and Phillips et al. (2010).

Terminal taxon ages were not precisely known for most extinct taxa but were approximated from fossil occurrence ages or geologic ages of formations in which taxa/ootaxa occur as reported by

Gill (2000), Rigby et al. (1993), Chiappe et al. (1998), Cunha et al. (2004) and Varricchio et al.

(2013).

3.3 RESULTS

3.3.1 Estimated D statistic

The estimated D statistic of nest type in the dataset of living archosaur species is -1.09, which is significantly different from 1 (p << 0.01) but not from 0 (p = 1.00). This D statistic indicates the presence of a strong phylogenetic bias in the distribution of nest types among 47

archosaurs.

3.3.2 Analysis of covariance (ANCOVA)

Eggshell porosity, relative to egg mass, was compared among living archosaurs with covered (n = 20) and open nest types (n = 107) (Figure 3.2). Eggshell porosity between the two types is normally distributed (p = 0.49 and 0.15 for open and covered nest types, respectively) and homogeneity of variances is observed (p = 0.83), indicating that a parametric test is appropriate for the dataset. Eggshell porosity is shown to be strongly correlated to egg mass in taxa with open nests (r = 0.87, p < 0.01) and moderately correlated in species with covered nests (r = 0.52, p <

0.05) (Table 3.4). Both conventional and phylogenetically-corrected ANCOVA reveal that the slopes between these two nest types are not significantly different (p >> 0.05). Furthermore, the intercept of the regressions, and thus eggshell porosity relative to egg mass, is found to be significantly higher in the covered nest type than the open nest type (p < 0.01, Table 3.5; Figure

3.2) except using the PGLS model where branch length was estimated from divergence time (p =

0.11), which showed no significant difference in intercept. Of all the conventional and phylogenetically-corrected methods used, the RegOU model, where divergence time was used for branch length assignment, has the lowest AIC value and is thus considered the best-fit regression model tested (Table 3.5). The AIC value of the PGLS model is much higher than for the other models (i.e., OLS and RegOU) regardless of the methods for branch length assignment, indicating that the PGLS models were the poorer fit for the dataset in this study.

3.3.3 Discriminant analysis

When phylogenetic relationships are not taken into consideration, the linear discriminant analysis reveals that nest type can be predicted from eggshell porosity and egg mass among living 48

archosaurs. From the dataset, 123 of the 127 extant bird and crocodylian species were classified correctly, resulting in an overall misclassification rate of only 3.15% (one open nester and three covered nesters were misclassified; Table 3.6). Applying this method to extinct archosaurs, crocodylomorphs (Bauruoolithus), possible ornithischians (Cairanoolithidae), sauropods

(Megaloolithidae), and two non-avian theropods (Lourinhanosaurus and Continuoolithus) were classified as covered nesters, whereas most oviraptorosaurs (Elongatoolithidae), Troodon (i.e.,

Prismatoolithus levis), and other avian and non-avian theropods (most Prismatoolithidae and moas) were classified as open nesters (Table 3.7; Figure 3.3). Unlike other elongatoolithid and prismatoolithid eggs, Elongatoolithus elongatus and Protoceratopsidovum minimum were classified into the covered nest type. Posterior probabilities of extinct taxa/ootaxa were generally high (> 0.70), indicating that their predicted nest types were well differentiated from the other types. Two ootaxa, E. elongatus and Pro. fluxuosum, have posterior probabilities close to 0.50, which indicates that their eggshell porosity is close to the threshold between covered and open nest types.

When phylogenetically-corrected methods are used, the pFDA reveals that the optimum

Pagel's lambda value is 0.56, which indicates that a moderately high phylogenetic bias exists in the dataset. The pFDA correctly classified 107 of 127 living species, resulting in an overall misclassification rate of 15.75% (Table 3.6). Although all extant archosaurs with an open nest type were classified correctly, none of the 20 species with a covered nest type was classified correctly.

Thus, the pFDA misclassification rate for extant covered nesters is 100%. The overall misclassification rate increases with increasing Pagal's lambda value, where the lowest misclassification rate was found at lambda values ≤0.01 (Figure 3.4). At optimal Pagel's lambda value, results of the phylogenetically-corrected discriminant analysis are consistent with the results of the conventional method (i.e., titanosaurs and Lourinhanosaurus are covered nesters, 49

and oviraptorosaurs, Troodon, and moas are open nesters; Table 3.8: Figure 3.5). Results of the pFDA do not change at non-optimal Pagel's lambda values except for oviraptorosaurs, which change to covered nesters at lambda values between 0.08 and 0.52, and for titanosaurs and

Lourinhanosaurus, which change to open nesters when lambda values approach one (Figure 3.6).

3.4 INTERPRETATION OF STATISTICAL RESULTS

Conventional and phylogenetically-corrected discriminant analyses produce results that differ for living taxa (i.e., higher misclassification rate in phylogenetically-corrected analysis, particularly among covered nesters) but that are consistent for fossil taxa (i.e., extinct taxa are assigned to the same nest type in both types of analysis). The LDA reveals that eggshell porosity

(relative to egg mass) predicts accurately nest type in living archosaurs (i.e., low misclassification rate). Since the precise taxonomic affinity, and hence phylogenetic position, is unknown for most fossil eggs, conventional methods such as LDA should be used for these specimens in order to obtain reliable inferences of nesting habits. In contrast when phylogenetic relationships are taken into consideration, misclassification rate is high among living species due to the presence of a phylogenetic bias in the dataset (i.e., high optimal Pagal’s lambda value in pFDA and low D statistic value). This phylogenetic bias is due to the fact that covered nests are restricted to two relatively basal clades of extant archosaurs (i.e., Crocodylia and Megapodiidae), resulting in a clumped phylogenetic distribution of this trait (Appendix 3.8). Previous studies that have used the pFDA method usually obtained lower optimal lambda values (< 0.20; see Schmitz and Motani,

2011; Hall et al., 2012; Close and Rayfield, 2012; Angielczyk and Schmitz, 2014), which indicates that the phylogenetic distribution of their traits was more randomly distributed (i.e., less clumped) than in this study.

When phylogenetic relationships were taken into consideration with the pFDA method, 50

the misclassification rates increased. The overall misclassification rate increases with increasing lambda values (Figure 3.4): the misclassification rate is at its lowest (3.15%) at low Pagal’s lambda values (≤0.01) and at its highest (15.75%) at high lambda values (≥0.55), including at the optimal lambda value. This pattern is opposite that observed by Motani and Schmitz (2011), the only study to have shown a change in misclassification rate with changing lambda values, where they observed the lowest misclassification rate around the optimal Pagal’s lambda value. The fact that pFDA assumes the Brownian motion model, which was the worst fit model for the dataset according to pcANCOVA, could explain the poor performance of this method in this study. Other methods, such as the Ornstein-Uhlenbeck model, which produced the best fit in pcANCOVA, may work better for the dataset, although this model cannot be performed in pFDA. A possible solution to this problem is to use LDA rather than pFDA until the latter method is developed further.

Regardless of the high misclassification rate of the pFDA, inferences of nest type for extinct archosaurs are consistent between LDA and pFDA (at optimal lambda value).

3.5 DISCUSSION

Results here reveal that eggshell porosity, expressed relative to egg mass, is highly correlated with nest type among living archosaurs in that eggs incubated in covered nests have a significantly higher eggshell porosity than those incubated in open nests. This newly discovered correlation permits the use of a discriminant analysis (LDA, pFDA) to infer nest type among extinct archosaurs, which could not be achieved with previous methods. Although pFDA struggled to correctly classify living species based on eggshell porosity, possibly due to built-in assumptions of evolutionary mode, results of LDA and pFDA were consistent for fossil taxa.

The eggshell porosity approach developed here is methodologically consistent and uses statistical rigor to infer nest type in extinct archosaurs, unlike the previous method based on water 51

vapor conductance (GH2O). For the GH2O method, nest type inference relies on comparisons between GH2O values measured experimentally for fresh eggs for living taxa and GH2O values estimated from egg/eggshell morphometric data for fossil taxa. The issues with this method are that: 1) the experimental and morphometric approaches do not produce results that are mutually consistent for living species, and thus comparisons between them should be avoided (Tanaka and

Zelenitsky, 2014a), and 2) a correlation between GH2O (neither experimental nor morphometric) and nest type has never been established in living taxa. In contrast, the eggshell porosity method proposed here relies exclusively on egg/eggshell morphometric data obtained from both living and extinct taxa and is based on a demonstrated (and statistically-significant) correlation between eggshell porosity and nest type.

This study reveals that sauropods, the theropod Lourinhanosaurus, and the potential ornithischian ootaxon Cairanoolithus had covered nests based on relatively high eggshell porosity, a result that is in agreement with most previous GH2O studies (Williams et al., 1984; Antunes et al.,

1998; Deeming, 2006; Grellet-Tinner et al., 2012), except one (Jackson et al., 2008). Most oviraptorosaurs (Elongatoolithidae) are classified here as open nesters due to relatively low porosity values (also determined by Wiemann et al., 2015), although several previous GH2O studies have inferred covered nests (e.g., Mou, 1992; Deeming, 2002c, 2006; Zhao et al., 2013). This discrepancy may be due to the fact that nest type was determined subjectively in these earlier studies, due to a lack of rigorous statistical analysis (Figure 3.3). Troodon formosus and two of three Protoceratopsidovum oospecies have relatively low eggshell porosity values and are inferred to have been open nesters, results consistent with the previously reported low GH2O values for

Troodon and Protoceratopsidovum (Sabath, 1991; Deeming, 2006; Varricchio et al., 2013).

When considered in a phylogenetic context, the results here shed light on the evolution of nest types among dinosaurs (Figure 3.7). The presence of covered nests in crocodylomorphs, 52

titanosaurs, the theropod Lourinhanosaurus, and probably ornithischians (Cairanoolithus) indicates that these nests were likely the primitive condition in Dinosauria and possibly

Archosauria. In contrast, open nests with partly or fully exposed eggs were present among oviraptorosaurs, troodontids, and birds, and thus were probably also present in the last common ancestor of oviraptorosaurs and troodontids (i.e., a non-avian maniraptoran). Open nests may have appeared even earlier in theropod evolution but a large phylogenetic gap in the fossil record of their eggs precludes a more precise determination (Figure 3.7). Nevertheless, results in the present study reveal that open nests first appeared in non-avian theropods well before the origin of Aves.

The evolutionary transition in nest types observed among non-avian theropods may also be linked to changes in other nesting habits, such as brooding behavior and the arrangement of eggs in the nest. A plausible scenario is that open nests and brooding behaviors evolved in association because the transfer of body heat to the eggs would be effective only if the eggs were exposed (at least partially) in the nest so as to be in contact with the parent (e.g., Varricchio et al.,

1997). This hypothesis is supported by the discovery of both oviraptorosaur and troodontid skeletons sitting atop or in contact with the eggs (Norell et al., 1995; Dong and Currie, 1996;

Varricchio et al., 1997 ; Clark et al., 1999; Fanti et al., 2012), which suggests that these early open nesters could have been brooders. Primitively, eggs laid in open nests may have been partially buried in substrate or nesting material, as suggested by the taphonomy of troodontid and enantiornithine bird clutches (Varricchio et al., 1999; Kurochkin et al., 2013; Varricchio and Berta,

2015) and the multi-layered arrangement of oviraptorosaur clutches (Wiemann et al., 2015). It is only later in avian evolution, presumably among euornithine birds, that eggs were left fully exposed in open nests, a condition observed in most extant brooding birds. Some neornithine taxa

(e.g., waders, , some waterfowl, , and ), however, likely secondarily evolved behaviors to partially bury their eggs during incubation (MacLean, 1974; Grant, 1982), 53

for either thermoregulation or concealment purposes (MacLean, 1974; Howell, 1979; Figure 3.7).

The evolution of open nests and brooding behavior may have played a key role in allowing maniraptoran theropods, including birds, to exploit a greater diversity of locations for nesting. Nest location for covered nesters (i.e., crocodylians and megapodes) is restricted to the ground because heat and humidity is required from the nesting materials/substrate for incubation

(Coombs, 1989). Conversely, reliance on body heat for egg incubation in fully open nesters probably freed maniraptorans to exploit new environments to build their nests (e.g., trees, cliffs, caves). Furthermore, this greater nesting freedom may have lessened the odds of nesting failure due to predation, flooding, or torrential rainfall, factors commonly adversely affecting the hatching success of covered nests on the ground (e.g., Brickhill, 1986; Coombs, 1989; Mazzotti, 1989;

VillamarIn-Jurado and Suarez, 2007), and consequently may have played a role in the evolutionary success and adaptive radiation of maniraptorans (Benson et al., 2014).

54

−1 Table 3.1. List of extinct archosaur taxa/ootaxa with estimated egg mass (M) and eggshell porosity (Ap∙Ls ) used in this study.

In order to supplement the data of Mou (1992), eggshell thickness was taken from Zhao (1975) for ootaxa that were indicated with *, because Mou (1992) provided only the thickness of the continuous layer as pore length.

−1 Family or oofamily Taxon or ootaxon log M log Ap∙Ls References

(Possible taxonomic affinity)

Krokolithidae (Crocodylomorph) Bauruoolithus fragilis 1.599 1.687 Oliveira et al. (2011)

Cairanoolithidae (Ornithischia?)/ Cairanoolithus dughii 3.468 2.732 Williams et al. (1984); Garcia and

Fusioolithidae Vianey-Liaud (2001)

Cairanoolithus 3.430 2.761 Garcia and Vianey-Liaud (2001)

roussetensis

Megaloolithidae (Sauropod)/ Megaloolithus aureliensis 3.705 3.327 Garcia and Vianey-Liaud (2001)

Fusioolithidae

Megaloolithus mammilare 3.716 3.055 Williams et al. (1984); Garcia and

Vianey-Liaud (2001)

Megaloolithus 3.351 2.602 Garcia and Vianey-Liaud (2001)

microtuberculata

Megaloolithus 3.107 2.703 Jackson et al. (2008); Grellet-Tinner et 55

−1 Family or oofamily Taxon or ootaxon log M log Ap∙Ls References

(Possible taxonomic affinity)

patagonicus/ titanosaur al. (2012)

sauropod

Megaloolithus petralta 3.420 2.788 Garcia and Vianey-Liaud (2001)

Megaloolithus 3.550 2.842 Garcia and Vianey-Liaud (2001)

pseudomamillare

Megaloolithus siruguei 3.622 3.327 Williams et al. (1984);

Lopez-Martinez et al. (2000); Garcia

and Vianey-Liaud (2001); Deeming

(2006); Jackson et al. (2008)

Megaloolithus cf. siruguei 3.325 3.138 Grigorescu et al. (1994, 2010)

Megaloolithus sp. 3.267 2.548 Zelenitsky pers obs. (cited in Deeming,

(recrystallized) 2006)

Megaloolithus sp. 3.430 3.090 Zelenitsky pers obs. (cited in Deeming,

(non-recrystallized) 2006)

Unknown megaloolithid 3.235 3.085 Williams et al. (1984) 56

−1 Family or oofamily Taxon or ootaxon log M log Ap∙Ls References

(Possible taxonomic affinity)

oospecies 1

Unknown megaloolithid 3.081 2.458 Grellet-Tinner et al. (2012)

oospecies 2

Oofamily Indet. (Non-avian Continuoolithus canadensis 2.320 1.785 Jackson et al. (2015) theropod)

Allosauroidea?/ ? Lourinhanosaurus antunesi 2.799 2.377 Antunes et al. (1998); Deeming (2006)

Elongatoolithidae Elongatoolithus andrewsi* 2.584 1.621 Zhao (1975); Mou (1992)

(Oviraptorosauria)

Elongatoolithus elongatus 2.411 1.659 Zhao et al. (2013)

Macroelongatoolithus 3.488 2.415 Zelenitsky pers obs. (cited in Deeming,

xixiaensis 2006)

Macroolithus rugustus* 2.772 1.642 Zhao (1975); Mou (1992)

Macroolithus yaotunensis*/ 2.911 1.835 Zhao (1975); Mou (1992); Wiemann et

oviraptorosaurs al. (2015)

Prismatoolithidae (Non-avian Prismatoolithus levis/ 2.463 1.213 Zelenitsky and Hills (1996); Varricchio 57

−1 Family or oofamily Taxon or ootaxon log M log Ap∙Ls References

(Possible taxonomic affinity) maniraptoran) Troodon formosus et al. (2013); Zelenitsky pers obs. (cited

in Deeming, 2006); this study

Protoceratopsidovum 2.411 1.602 Ornamented protoceratopsid egg in

fluxuosum Sabath (1991)

Protoceratopsidovum 2.106 1.523 Thin-shelled protoceratopsid egg in

minimum Sabath (1991)

Protoceratopsidovum 2.380 1.465 Smooth-shelled protoceratopsid egg in

sincerum Sabath (1991)

Sankofa pyrenaica 1.788 0.478 Lopez-Martinez and Vicens (2012)

Dinornithiformes Euryapteryx sp. 2.771 1.765 Gill (2006); Huynen et al. (2010); this

study

Pachyornis geranoides 2.771 1.41 Gill (2006); Huynen et al. (2010); this

study

58

Table 3.2. List of variables used for this study, modified from Tanaka and Zelenitsky

(2014a).

Variable Definition Unit

A Individual pore area (μm2)

2 Ap Total pore area (mm )

−1 Ap∙Ls Eggshell porosity (mm)

2 As Surface area of eggshell (mm )

B Maximum egg breadth (mm)

D Pore density (mm−2)

L Maximum egg length (mm)

Ls Shell thickness (= pore length) (mm)

M Egg mass (g)

N Total number of pores (per egg)

V Egg volume (mm3)

59

Table 3.3. List of equations used for this study, modified from Tanaka and Zelenitsky

(2014a).

Equation Sources

Ap = A∙As∙D = A∙N Seymour (1979)

0.666 As = 4.951∙V Paganelli et al. (1974)

M = 5.48∙10-4∙L∙B2 Hoyt (1979)

V = 0.51∙L∙B2 Hoyt (1979)

60

Table 3.4. Results of conventional OLS regression models for living archosaur species.

Abbreviations: CI, 95% confidence interval; n, sample size; r2, coefficient of determination.

Type n Slope CI of slope Intercept CI of intercept r2

Covered nester 20 0.874 0.161 to 1.587 -0.290 -1.699 to 1.119 0.269

Open nester 107 1.117 0.994 to 1.239 -1.453 -1.668 to -1.239 0.756

61

Table 3.5. Results of conventional and phylogenetically-corrected ANCOVA for living archosaur species.

Abbreviations: AIC, akaike information criterion; d.f., degree of freedom; F, test statistic; OLS, ordinary least-squares; PGLS, phylogenetically-generalized least-squares assumed Brownian motion process; RegOU, phylogenetic regression with Ornstein-Uhlenbeck process. The lowest

AIC value is shown with bold.

Branch length assignment Model F d.f. p AIC

None OLS 108.797 1, 124 << 0.01 25.580

Branch length = 1 PGLS 11.568 1, 124 0.001 53.669

Branch length = 1 RegOU 70.111 1, 124 << 0.01 25.007

Divergence time PGLS 2.542 1, 124 0.113 72.841

Divergence time RegOU 81.941 1, 124 << 0.01 17.578

62

Table 3.6. Cross-classification/confusion matrix from LDA and pFDA.

The true classifications are along the top and the predicted classifications are on the left-hand side.

LDA pFDA

Covered nest Open nest Covered nest Open nest

Covered nest 17 1 0 0

Open nest 3 106 20 107

% Correct 85.000% 99.065% 0% 100%

Overall 3.150% 15.748%

misclassification rate

63

Table 3.7. Inferred nest types for extinct archosaurs based on the linear discriminant analysis.

Family/oofamily Taxon/ootaxon Prediction Posterior probabilities

Covered nest Open nest

Krokolithidae Bauruoolithus fragilis Covered nest 1.000 0.000

Cairanoolithidae/ Fusioolithidae Cairanoolithus dughii Covered nest 0.747 0.253

Cairanoolithus roussetensis Covered nest 0.849 0.151

Megaloolithidae/ Fusioolithidae Megaloolithus aureliensis Covered nest 0.992 0.008

Megaloolithus mammilare Covered nest 0.881 0.119

Megaloolithus microtuberculata Covered nest 0.704 0.296

Megaloolithus patagonicus Covered nest 0.985 0.015

(titanosaurs)

Megaloolithus petralta Covered nest 0.889 0.111

Megaloolithus pseudomamillare Covered nest 0.804 0.196

Megaloolithus siruguei Covered nest 0.996 0.004

Megaloolithus cf. siruguei Covered nest 0.999 0.001

Megaloolithus sp. Covered nest 0.753 0.247

(recrystallized) 64

Family/oofamily Taxon/ootaxon Prediction Posterior probabilities

Covered nest Open nest

Megaloolithus sp. Covered nest 0.994 0.006

(non-recrystallized)

Unknown megaloolithid Covered nest 0.896 0.104

oospecies 1

Unknown megaloolithid Covered nest 0.999 0.001

oospecies 2

Oofamily Indet. Continuoolithus canadensis Covered nest 0.909 0.091

Allosauroidea?/ Coelurosauria? Lourinhanosaurus antunesi Covered nest 0.978 0.022

Elongatoolithidae Elongatoolithus andrewsi Open nest 0.139 0.861

Elongatoolithus elongatus Covered nest 0.543 0.457

Macroelongatoolithus xixiaensis Open nest 0.091 0.909

Macroolithus rugustus Open nest 0.033 0.967

Macroolithus yaotunensis Open nest 0.061 0.939

(oviraptorosaurs)

Prismatoolithidae Prismatoolithus levis (Troodon Open nest 0.008 0.992 65

Family/oofamily Taxon/ootaxon Prediction Posterior probabilities

Covered nest Open nest

formosus)

Protoceratopsidovum fluxuosum Open nest 0.402 0.598

Protoceratopsidovum minimum Covered nest 0.842 0.158

Protoceratopsidovum sincerum Open nest 0.185 0.815

Sankofa pyrenaica Open nest 0.003 0.997

Dinornithidae Euryapteryx sp. Open nest 0.107 0.893

Pachyornis geranoides Open nest 0.003 0.997

66

Table 3.8. Inferred dinosaur nest types based on the phylogenetic flexible discriminant analysis. Taxon/ootaxon Prediction Posterior probabilities Covered nest Open nest Megaloolithus patagonicus (titanosaurs) Covered nest 0.888 0.112 Lourinhanosaurus antunesi Covered nest 0.898 0.102 Macroolithus yaotunensis Open nest 0.487 0.513 (oviraptorosaurs) Prismatoolithus levis (Troodon formosus) Open nest 0.357 0.643 Euryapteryx sp. Open nest 0.358 0.642 Pachyornis geranoides Open nest 0.243 0.757

67

Figure 3.1. Porosity of archosaur eggshell. Schematic diagram of archosaur eggshell with high porosity (a) and low porosity (b), modified from Tanaka and Zelenitsky (2014a); tangential thin sections of living covered nester Caiman latirostris (c), living open nester Pavo cristatus (d), and non-avian maniraptoran Troodon formosus (e). Abbreviations: A, individual pore area; D, pore density; Ls, pore length. Arrows indicate pore canals.

68

Figure 3.2. Bivariate plot of eggshell porosity and egg mass between living covered and open nesters. Eggshell porosity relative to egg mass is highly correlated to nest types (p < 0.01), as reflected by the different regression models between closed and open nesters.

69

Figure 3.3. Bivariate plot of eggshell porosity and egg mass in both living and extinct archosaur taxa/ootaxa. Titanosaurs and Lourinhanosaurus show high eggshell porosity, comparable to living species with covered nests. In contrast, oviraptorosaurs, Troodon, and moas show lower eggshell porosity, similar to species with open nests.

70

Figure 3.4. Misclassification rate of pFDA for living species through changing Pagel's lambda values. Dash line shows the optimal lambda value of 0.56. Note that the overall misclassification rate increases with increasing lambda values from 0 to 1.

71

Figure 3.5. Comparison of the discriminant function between covered and open nesters in living and fossil archosaurs. Horizontal bars inside boxes represent medians, lower and upper ends of boxes are the 25% and 75% quartiles, respectively, and whiskers represent the smallest and largest cases. Outliers are represented by dots and extremes by diamonds. Note that covered nesters show relatively lower values than open nesters.

72

Figure 3.6. Inferred nest type for six extinct archosaurs as a function of Pagel's lambda values. Inferred nest type is generally consistent across all lambda values, except for oviraptorosaurs where inferred nest type changes when the lambda value varies between 0.08 and 0.52, and for titanosaurs and Lourinhanosaurus, which change to open nesters when lambda values approach one. The yellow line indicates the optimal lambda value (0.56).

73

Figure 3.7. Evolution of nest types among archosaurs. (a) Phylogeny of archosaurs with inferred nest types based on eggshell porosity and taphonomic evidence. Covered nests are the primitive condition for dinosaurs; open nests and brooding behavior were present among non-avian maniraptoran theropods but may have first appeared earlier. Although the eggs of early open nesters were still partially covered by substrate, open nests with fully exposed eggs likely arose among Euornithes. (b) Phylogeny of Neornithes with inferred nest types based on eggshell porosity (Emeidae) and literature (other birds). Open nests with fully exposed eggs are the primitive condition for modern birds, although secondary reversal to partial egg burial occurred independently in several clades. Information for bird orders which include species that partially bury the eggs (Charadriiformes) or occasionally cover the eggs in open nests (, Anseriformes, Charadriiformes, , Passeriformes, Podicipediformes, Struthioniformes, Tinamiformes) was taken from Broley (1947), MacLean (1974), Marchant and Higgins (1990), White and Kennedy (1997), Davies (2002), Kear (2005), and Kathju (2007). Cladograms are based on Jarvis et al. (2014), Nesbitt (2011), Mitchell et al. (2014), Novas et al. (2015), and Wang et al. (2015).

74

75

CHAPTER 4: NEST MATERIAL, NEST TYPE, AND INCUBATION HEAT SOURCES IN DINOSAURS: IMPLICATIONS FOR DINOSAUR NESTING AT HIGH LATITUDE

4.1 INTRODUCTION Living archosaur species nest in various locations and environments. Nesting sites of brooding bird species range from on water to island cliffs, from desert to montane regions, or from equator to polar regions (Carey, 2002; Hansell and Deeming, 2002). Brooding behavior possibly enabled them to nest in various environments because adult contact provides constant, adequate heat for incubation of eggs, regardless of climatic zones at nesting sites (see Tanaka et al., 2015). While brooding keeps eggs warm in cold climate (e.g., Emperor : Le Maho, 1977), this behavior also provides a shade for eggs in hot deserts in order to avoid lethal egg temperature (e.g., Gray : Howell et al., 1974). In contrast, species that cover their eggs with nest materials or substrate (i.e., mound or in-filled hole nests) (Booth and Jones, 2002; Brazaitis and Watanabe, 2011) obtain incubation heat from the surrounding environments (e.g., microbial respiration of decaying plant materials, geothermal activity, and solar radiation: Ferguson, 1985; Magnusson et al., 1985; Booth and Jones, 2002). As such, the geographic distribution of extant covered nesters (i.e., crocodylians and megapode birds) is generally limited to low latitudes, up to 35° from the equator in crocodylians (Lance, 2003) and up to 37° from the equator in megapodes (Jones et al., 1995). Distribution of nesting sites thus could be related, in part, to nesting methods (incubation heat sources and nest structure) in archosaurs. Fossil dinosaur eggs, eggshells, and nests are found around the world, including Asia, Europe, Africa, North America, and South America (Carpenter, 1999). Although major dinosaur egg localities are often located at mid latitudes (e.g., , southwestern Canada, , , , , Portugal, , South Africa, , , and USA: Carpenter, 1999; Huh and Zelenitsky, 2002; Reisz et al., 2012), dinosaur eggshells have been reported from the high Arctic (Godefroit et al., 2009). In addition, dinosaur eggs and eggshells are recovered from various sediments (e.g., siliciclastic mudstone, siltstone, sandstone, and conglomerate as well as carbonate rocks, with or without pedogenic features: e.g., Khosla and Sahni, 1995; Varricchio et al., 1999; Liang et al., 2009) and depositional environments (e.g.,

76

fluvial, lacustrine, littoral, and eolian: Paik et al., 2012). The distribution of dinosaur eggs and eggshells thus casts an intriguing question related to how dinosaurs incubated their eggs in various environments and regions, including high arctic regions. However, a relationship between incubation method and paleogeography of nesting sites remains poorly understood in dinosaurs. Research on incubation heat sources is limited in dinosaurs (Grellet-Tinner and Fiorelli, 2010) because nest materials and structures are rarely preserved in the fossil record (Chiappe et al., 2004; Zelenitsky and Therrien, 2008b). Studies on nest taphonomy and eggshell porosity indicate that oviraptorosaurs and troodontids built at least partially open nests and practiced a brooding-like behavior (Dong and Currie, 1996; Clark et al., 1999; Fanti et al., 2012; Varricchio et al., 2013; Tanaka et al., 2015; Wiemann et al., 2015); as such, these species would theoretically have been adapted for nesting at high latitudes. In contrast, nesting in cool or cold environments is presumably more challenging for dinosaur species that are inferred to have built covered nests (e.g., ornithischians, sauropodomorphs, and Lourinhanosaurus; Antunes et al., 1998; Deeming, 2006; Grellet-Tinner et al., 2012; Tanaka et al., 2015) because incubation heat is obtained from the surrounding environments. Interestingly, hadrosaur and sauropod eggs and eggshells are known from a wide variety of paleolatitudes, reaching up to 76.7°N (i.e., hadrosaurs of the Kakanaut Formation in northernmost Siberia; Godefroit et al., 2009). Although the Late Cretaceous Arctic climate was admittedly generally warmer than today (e.g., Tarduno et al., 1998; Jenkyns et al., 2004), estimated mean annual and summer temperature in the Kakanaut area during the Maastrichtian were around 10°C and 19°C, respectively (Golovneva, 2000; Herman et al., 2016). An understanding of the types/structures of nests and incubation heat sources used by dinosaurs is thus essential to reconstruct the nesting methods of dinosaurs in various environments, including the high Arctic.

4.2 RESEARCH APPROACHES In order to better understand the incubation methods of dinosaurs, in particular those of covered nesters, at various latitudes, including polar regions, relationships are investigated between nests and nesting environments among both extant and fossil archosaur taxa (see Figure 4.1 for the summary of tests). First, it is explored whether ambient air temperature at the nesting sites affects the incubation temperature in covered nests built by living crocodylians and megapodes. This is done by subdividing nests into two categories based on the source of

77

incubation heat used: 1) nests that derive incubation heat mainly from decomposing organic matter produced by microbial activity, and 2) nests that derive incubation heat from solar radiation. If nest temperature is unrelated to ambient air temperature in either group, then these nests could theoretically be used at high latitudes. Secondly, it is explored if incubation heat sources and/or nest structure (i.e., mounds and in-filled holes) of living covered nesters can be inferred from nest materials/substrates. Because nesting material/substrate and nest structure play a significant role in the generation and/or transfer of heat to the eggs in covered nesters (Ferguson, 1985; Magnusson et al., 1985; Booth and Jones, 2002), it can be hypothesized that the type of nest materials/substrates (e.g., plant materials, , soils: Joanen, 1969; Lutz and Dunbar-Cooper, 1984; Jones et al., 1995; Somaweera and Shine, 2013) used in building nests is correlated to incubation heat sources and/or types of nest structures. For this, lithologic data on fossil nests, eggs, and eggshells for various dinosaur ootaxa were gleaned from the literature and statistically compared to determine if certain dinosaurs favored certain types of sediments to build their nests. If a relationship between these variables can be established, then it will be possible to infer incubation heat sources and nest structures of dinosaurs based on the lithology of their nests.

4.3 MATERIAL AND METHODS 4.3.1 Datasets and statistical tests for extant taxa Mean nest and ambient air temperature for Crocodylia and Megapodiidae was obtained from the literature and the difference between these two variables was calculated for each species/population (Appendix 4.1). Nest temperature was measured at the center of the nest or around the egg chamber during incubation. Ambient air temperature represents either measurements taken around the nests or averages between minimum and maximum air temperature during nesting season. Ambient air temperature at crocodylian and megapode nesting sites are fairly stable; fluctuations of the temperature during nesting season and the mean daily temperature are less than 10°C in most cases. Heat produced by embryonic metabolism may contribute to nest temperature (e.g., Modha, 1967; Webb et al., 1983b; Magnusson et al., 1985; Seymour et al., 1987); however, metabolic heat was not considered in this study because it is difficult to distinguish metabolic heat from nest temperature. Embryonic metabolic heat can potentially raise egg temperature 1–3 °C in crocodylian and megapode nests (Webb et al., 1983b; Seymour et al., 1987; Webb and Cooper-Preston, 1989), and could be a small contribution to nest

78

temperature (Staton and Dixon, 1977; Magnusson et al., 1985). Paired-samples T-tests were used to compare mean nest and air temperature for nests with incubation heat from microbial respiration and from solar radiation. All statistical analyses were conducted with IBM SPSS Statistics v. 22.0.0 (IBM SPSS Inc.) at an α level of 0.05. In order to determine if the type of nesting material/substrate differs between nests of different heat sources and structure (i.e., mound and in-filled hole nests), data on nest material/substrate and nest structure in 42 species of Crocodylia and Megapodiidae were gleaned from the literature (Appendix 4.2) and compared using Fisher's exact test. Incubation heat sources were classified as predominantly from either: 1) inorganic heat source (i.e., solar radiation and geothermal activities), or 2) organic heat source (i.e., microbial activity associated with decaying plant material or termite nest with potential supplemental contribution from inorganic heat sources). Nest material/substrate was classified as either: 1) sand-dominated, 2) clay-dominated, or 3) plant material- (e.g., leaves, twigs/sticks, roots, tree stump: Jones et al., 1995) and/or soil-dominated. Nest structure was classified as either a mound (pile of nest materials on the burrowed or flat ground) or in-filled hole nest (underground hole).

4.3.2 Datasets and statistical tests for fossil taxa/ootaxa In order to test whether dinosaur nests are associated with particular types of sediments, lithologic data of dinosaur egg specimens were obtained from the literature supplemented with museum specimen observations (Table 4.1). For this study, two datasets were compiled based on the quality of fossil preservation: 1) clutches and nesting grounds in which the eggs are presumably in-situ (Appendix 4.3); and 2) isolated eggs and eggshells potentially not in-situ (assumedly ex-situ) (Appendix 4.4). Both datasets consist of various dinosaur ootaxa, including eggs that can be accurately referred to specific dinosaur taxa (Table 4.2). In this study, oofamily Spheroolithidae is referred to as hadrosaur eggs, Megaloolithidae and Faveoloolithidae as sauropodomorph eggs, Elongatoolithidae as oviraptorosaur eggs, and Prismatoolithidae as troodontid (and possibly other non-oviraptorosaur maniraptoran) eggs. Preprismatoolithus eggs were excluded from the oofamily Prismatoolithidae because they belong to Allosaurus and Lourinhanosaurus, which are not maniraptorans (Mateus et al., 2006; Carrano et al., 2013). In the in-situ dataset, multiple clutches from a single ootaxon discovered within a single horizon (including colonial grounds) were considered as one occurrence. Clutches of a single ootaxon

79

discovered from multiple horizons within a single locality were considered as multiple occurrences by counting the number of clutch horizons. Lithologic data were classified into either fine-grained (i.e., siltstone and mudstone, including ) or coarse-grained (i.e., sandstone and conglomerate). One-way χ2-tests were conducted in order to test whether a particular type of sediments (i.e., fine-grained and coarse-grained) is associated with each ootaxon, following the method of Lyson and Longrich (2011). Evidence of paleosol and/or pedogenetic features (e.g., root traces, rhizoliths, caliche nodules, slickensides, , and pedotubules: Retallack, 2001) in clutch horizons was also recorded for the in-situ dataset. Although that developed after burial of the nests (e.g., Lopez-Martinez et al., 2000; Diaz-Molina et al., 2007; Grigorescu et al., 2010) were not incorporated into the dataset, it was often difficult to determine in pedogenic development occurred before or after nesting. Additionally, the paleogeographic distribution of egg/eggshell remains was investigated for major dinosaur oofamilies (Appendix 4.5). Information about fossil egg/eggshell localities from around the world was gathered from the literature, and their approximate paleolatitudes were calculated using the website paleolatitude.org (van Hinsbergen et al., 2015). If paleolatitudes could not be calculated with this website, the Paleobiology Database (https://paleobiodb.org/#/) was used.

4.4 RESULTS 4.4.1 Dataset for extant taxa Nest temperature derived from both microbial respiration and solar radiation is significantly different from ambient air temperature during incubation (t = 7.26, d.f. = 16, p < 0.01 for nests that use heat from microbial respiration and t = 3.93, d.f. = 8, p < 0.01 for nests that use heat from solar radiation). In all cases, except for one (i.e., Alligator mississippiensis: Joanen, 1969), nest temperature is higher than ambient air temperature. For nests that use incubation heat derived from solar radiation, the difference between nest and air temperature ranged from 0.98 to 6.75 °C (mean 3.93 °C: standard deviation 2.29), whereas the temperature difference was considerably higher (mean 7.26 °C) and more variable (from -2.50 to 22.20 °C: standard deviation 6.35) in nests with microbial respiration (Figure 4.2). Materials/substrates preferentially used to build nests are different depending on the

80

source of incubation heat (p < 0.05: Figure 4.3). In nests where inorganic heat sources are used for incubation (i.e., heat from solar radiation and geothermal activities), sand is a more common nesting substrate (n = 9) than plant materials and/or soils (n = 5). In contrast, nests that use organic heat sources (i.e., predominantly heat from microbial respiration of decaying plant materials and termite nests) for incubation are more commonly composed of plant materials and/or soils (n = 18) than sand (n = 6). Nests made of clay were not observed among the species considered (n = 0). Interestingly, mound and in-filled hole nests are shown to be built from different nest materials/substrates (p < 0.01: Figure 4.4). The majority of mound nesters use plant materials and/or soils (n = 33) for their nests, and use of sand (n = 9) or clay (n = 0) is much less common. However, in-filled hole nesters commonly lay eggs within sand (n = 12), whereas clay (n = 2) or plant materials (n = 6) are less commonly used.

4.4.2 Dataset for dinosaur taxa/ootaxa Overall, dinosaur egg/eggshell remains occur in both fine- and coarse-grained sediments about equally (1: 0.89 and 1: 0.81 in the in-situ and ex-situ datasets, respectively: Tables 4.3 and 4.4). However, there are regional variations of clutch occurrences in the in-situ dataset (Table 4.5). For example, sauropods laying clutches of megaloolithid eggs tend to be found from fine-grained pedogenic deposits in Romania and Spain but from coarse-grained pedogenic deposits in India. Clutches of maniraptorans (i.e., oviraptorosaurs and troodontids) are mainly recovered from fine-grained deposits in China (for oviraptorosaurs) and in North America (for troodontids) but from coarse-grained deposits in Mongolia. Faveoloolithid clutches are mainly found from coarse-grained deposits regardless of the regions of the egg localities. One-way χ2-tests for the in-situ dataset revealed that certain types of dinosaur egg clutches are not randomly distributed with respect to lithology (Table 4.3: Figure 4.5a). Faveoloolithid clutches are significantly associated with coarse-grained deposits, whereas megaloolithid clutches are associated with fine-grained deposits. Although hadrosaur clutches (Spheroolithidae) were not subjected to a statistical test due to the small sample size, they tend to be associated with fine-grained deposits (n = 11) rather than coarse-grained deposits (n = 3). In contrast, there was no significant lithological association observed for clutches of Dendroolithidae, oviraptorosaurs (Elongatoolithidae) and troodontids (Prismatoolithidae). For the ex-situ dataset, one-way χ2-tests revealed no clear relationships between lithology and any taxa/ootaxa (Figure 4.5b). No

81

significant correlation of lithology was detected in hadrosaurs, sauropods laying megaloolithids and faveoloolithids, oviraptorosaurs, and troodontids (Table 4.4). Dendroolithidae was not tested due to the small sample size (n = 4). Evidence of pedogenetic development is commonly found in clutch horizons of hadrosaurs and sauropods laying megaloolithids, with 64% and 68%, respectively (Figure 4.6). However, pedogenetic features were not common in other taxa/ootaxa, such as Dendroolithidae, oviraptorosaurs, one type of sauropod eggs (Faveoloolithidae), and troodontids, occurring in equal to or less than 30% of the clutches in each group. Estimated paleolatitude of dinosaur egg/eggshell localities indicates that the vast majority of localities occur between 20° and 60° of the paleoequator (Table 4.6: Figure 4.7). Localities outside this latitudinal range are uncommon, but one extremely high paleolatitude (76.7°N) locality in northeastern Russia has produced spheroolithid (hadrosaur) and prismatoolithid (troodontid) eggshells (Godefroit et al., 2009).

4.5 DISCUSSION This study reveals that archosaur taxa, both living and extinct, preferentially use specific materials/substrates related to their nesting behaviors. Species that build mound nests preferentially use plant materials or soil (e.g., leaves, twigs/sticks, roots, tree stump: Joanen, 1969; Jones et al., 1995) as nest materials, whereas species that build in-filled hole nests often lay eggs in sand. Incubation heat source is also associated with nest materials/substrates in that nests that rely on organic heat sources for incubation are often composed of plant materials and/or soil, whereas those that rely on inorganic heat sources are usually built in sand. Regardless of incubation heat source, nest temperature tends to be higher than ambient air temperature. Although nest temperature derived from solar radiation is only somewhat higher than ambient air temperature (mean difference 3.93 °C), nest temperature derived from organic heat sources (i.e., microbial activity) is much higher than air temperature (mean difference 7.26 °C). Also, nests relying on organic heat sources are found in a wider range of ambient air temperature than those using solar radiation, indicating that organic heat sources are used in more various environments than solar radiation. In non-avian dinosaurs, the dataset of in-situ clutches and nesting grounds suggests that certain dinosaurs favored certain types of sediments to build their nests (e.g., pedogenetic fine-grained sediments for hadrosaurs and sauropods laying megaloolithids and coarse-grained

82

sediments for sauropods laying faveoloolithids). Dinosaur eggshells and isolated eggs that are unlikely to be in-situ, however, are not correlated to a particular lithology, presumably reflecting the fact that some of them have been transported and deposited away from the original sediments.

4.5.1 Nesting methods of extant species For living mound and in-filled hole nesters, selection of nesting grounds is important for building a nest and acquiring incubation heat. Mound nesters primarily use plant materials or soil for their nests in order to obtain heat from microbial activity. As such, they generally build their nests in marshes, swamps, forests, or scrubs (e.g., Frith, 1956; Joanen, 1969; Webb et al., 1977; Cintra, 1988), where soils are well-developed and/or abundant plant material is available. In contrast, species that build in-filled hole nests often lay eggs in sand from sandbars, river banks, beaches, and volcanic fields, where vegetation is limited (e.g., Ripley, 1960; Falanruw, 1975; Bolhm, 1982; Thorbjarnarson and Hernandez, 1993). In-filled hole nesters can locate a suitable substrate and depth for incubation that provides stable heat by digging test holes (e.g., Somaweera and Shine, 2013). For species that use heat from solar radiation or geothermal activities, sand is probably a better substrate than mud because sand has a better thermal conductivity (Chabreck, 1973; Charruau et al., 2010; Charruau, 2012). Thus, species that build mound and in-filled hole nests preferentially select substrates for nesting. Organic heat sources (i.e., microbial respiration) are likely more suitable for nesting in cooler regions. This study demonstrates that microbial respiration provides adequate incubation heat even though ambient air temperature is much lower than nest temperature. The same conclusion has been obtained in field observations. For example, one study of a mound nester, Aepypodius (Brush-turkey), found that heat produced by microbial respiration rose nest temperature to 60 °C even though ambient air temperature was only 15.6 °C (Frith, 1956). Another example documented that the core mound temperature of Alectura lathami (Australian Brush-turkey) was maintained at 33 °C while the mean ambient air temperature was 18 °C (Seymour and Bradford, 1992). Thus, nest temperature derived from microbial respiration is not correlated with ambient air temperature (Jones, 1988; Sinclair, 2001; VillamarIn-Jurado and Suarez, 2007), and is significantly higher than air temperature (Chabreck, 1973; Magnusson, 1985; Cintra, 1988; Jones, 1988; Sinclair, 2001). These results suggest that mound nesters, by using microbial respiration, have the potential to nest in cooler regions or higher latitudes than

83

species using heat derived from solar radiation. In fact, it has been reported that captive mound-nesting megapodes naturally incubated at higher latitudes (i.e., Melbourne Zoo and Zoo Frankfurt: Fleay, 1937; Baltin, 1969) than the bird’s original habitats (Figure 4.8). Like microbial respiration, geothermal activity is a potential heat source for successful incubation in cooler regions because geothermal heat is independent from ambient air temperature. While geothermal heat is theoretically available anywhere in the world where geothermal activity exists, current distribution of species that use geothermal heat for incubation is limited to tropical oceanic islands (Figure 4.8). Only six of 22 species of megapodes (genera Megapodius and Macrocephalon) are known to use geothermal heat, although they also use other heat sources (i.e., microbial respiration and/or solar radiation) (MacKinnon, 1978; Jones et al., 1995; Harris et al., 2014). Due to the limitations of previous studies, further investigation is required to test the correlation between their nest and ambient air temperatures. Unlike microbial respiration and presumably geothermal activity, solar radiation is only available as a heat source when air temperature is close to nest temperature (Figure 4.2). Although in-filled hole nests warmed by solar radiation have a nest temperature that is slightly higher than ambient air temperature, previous studies have found that their temperature is correlated with ambient air temperature (Magnusson, 1979; Somaweera and Shine, 2013). As such, the distribution of in-filled hole nesters that derive incubation heat from solar radiation is primarily limited to within 30° of the equator (Figure 4.8). Thus, solar radiation cannot be used as a source of incubation heat in relatively cool or cold environments, unlike microbial respiration and potentially geothermal activity. Although incubation with in-filled hole nests is only utilized in environments where enough solar radiation or geothermal activity is available, this incubation style requires less effort than incubation with mounds (Jones et al., 1995). It is not necessary to collect nest materials, but species that build in-filled hole nests locate warm sand grounds and dig a hole where an optimal incubation temperature is obtained (e.g., Modha, 1967; Roper, 1983; Goth and Vogel, 1996–1997; Heij et al., 1997). Their nesting sites are thus limited at sandy substrates in warm environments (Jones et al., 1995). In contrast, mound nests using microbial respiration can generate adequate incubation heat even at higher latitudes, but this style requires a large amount of nest materials. In Crocodylus porosus (Saltwater ), for example, a mound can be 2–3 meters in diameter with a height up to one meter (Ferguson, 1985). A mound of Alectura lathami can be 3–4 meters in

84

diameter and over two tonnes in weight (Jones, 1988); they take 3–6 weeks for construction of a mound (Jones et al., 1995). Thus, there is likely a relationship between the range of nesting environments and parental effort in covered nesters. Bird species that build open nests are not tested in this study because incubation heat is directly transferred from brooding adults, no matter what nest materials are used. Their nest materials/substrates (e.g., plants, lichen, fungi, feathers, fur, silk, stones, or mud: Hansell, 2000) as well as nest locations (e.g., on ground, trees, cliffs, water, or in cavities: Hansell, 2000) are highly diverse. Nevertheless, adult contact incubation provides consistent incubation heat to the eggs, although the insulating effects of nest materials/substrates can assist regulation of nest temperature (e.g., MacLean, 1974; Davis et al., 1984; Thompson and Raveling, 1988; Prokop and Trnka, 2011). Due to their nesting style, brooding birds can nest in diverse locations and environments, including severe Antarctic and mountain regions (Watson, 1975; Carey, 2002), otherwise prohibitive for non-brooding birds.

4.5.2 Nest structures, nesting substrates, and incubation heat sources of extinct dinosaurs This study also provides insight into the nest structure and sources of incubation heat of dinosaurs (Figure 4.9). It is revealed that in-situ eggs of hadrosaurs and sauropods laying megaloolithid eggs are associated with fine-grained pedogenic sediments, indicating that many (but not necessarily all) hadrosaurs and sauropods laying megaloolithid eggs nested on grounds similar to many mound nesting megapodes and crocodylians. Although the lithologic type (i.e., fine-grained vs. coarse-grained) of megaloolithid clutches varies among localities (countries), pedogenic features are consistently associated with their clutches regardless of the localities, indicating that they favored a soil ground. Combined with the results of eggshell porosity studies indicative of covered nests (Tanaka et al., 2015), hadrosaurs and sauropods laying megaloolithid eggs are inferred to have built mound nests using soil and plant materials. Evidence of nest excavation in sauropods laying megaloolithid eggs (e.g., Mohabey, 1996; Chiappe et al., 2004; Vila et al., 2010a, 2010b) suggests that eggs were laid in depressions before being covered with plant material or soil, as seen in living mound nesters (Sinclair, 2001; Medem, 1981; Mazzotti, 1989), although evidence of fossil plant matter is only occasionally preserved (Lopez-Martinez et al., 2000; Chiappe et al., 2004; Vila et al., 2010a). Consequently, eggs of hadrosaurs and sauropods laying megaloolithid eggs were likely incubated mainly with organic heat sources, by the

85

decomposition of plant matter in the mound. The eggs of other sauropods, those laying faveoloolithid eggs, are also inferred to have been incubated in covered nests based on eggshell porosity studies (Deeming, 2006; Grellet-Tinner et al., 2012), but their strong association with non-pedogenic, coarse-grained sediments (particularly sandstone) suggests they were incubated in in-filled hole nests rather than mounds. As such, these sauropods presumably incubated their eggs using inorganic heat sources, like solar radiation or, in at least one case, geothermal heat (Grellet-Tinner and Fiorelli, 2010). Eggs of Lourinhanosaurus and possibly Allosaurus (oogenus Preprismatoolithus) were also likely incubated in covered nests based on the studies of eggshell porosity (Antunes et al., 1998; Tanaka et al., 2015); however, their incubation heat source cannot be inferred due to the limited samples (n = 2). In contrast to hadrosaur and sauropod nests, the nests of oviraptorosaurs and troodontids are not associated with any specific lithology, indicating that they may not have derived incubation heat from nesting materials. Several lines of evidence indeed indicate that they built at least partially open nests in which the eggs were partially exposed (Varricchio et al., 2013; Tanaka et al., 2015; Wiemann et al., 2015) and brooded their eggs (Dong and Currie, 1996; Varricchio et al., 1997; Clark et al., 1999; Fanti et al., 2012). As such, these theropods were probably more versatile and thus may have been less selective in choosing a substrate for nesting as there was no reliance on external heat sources.

4.5.3 Distribution of dinosaur eggs and implications for dinosaur nesting at high latitude Knowledge of the sources of incubation heat for dinosaurs provides insight into the paleogeographic distribution of dinosaur egg remains (Figures 4.7, 4.10 and 4.11). The majority of dinosaur egg/eggshell localities occur between 20° and 60° of the paleoequator, where mean sea-surface temperature was around 20°C to over 30°C in Turonian (Takashima et al., 2006), but the distributions differ among ootaxa (which represent egg-laying taxa). For example, the paleogeographic distributions of sauropod (megaloolithids and faveoloolithids) and oviraptorosaur eggs/eggshells are limited to mid-paleolatitudes (18.3°–49.0° from the equator). In contrast, hadrosaurs and some maniraptoran theropods (prismatoolithids) show broader paleogeographic distributions, including mid- through high paleolatitudes (20.0°–76.7° from the equator) and across various climatic zones. Limited reports of dinosaur eggs/eggshells from low and high paleolatitudes (< 20° and > 60° from the equator) could be related to the fact that dinosaur fossil localities are much less common in these latitudes than in mid-latitudes

86

(Weishampel et al., 2004). Also, some dinosaur groups may not have nested in high latitudes. These differences in paleolatitudinal distribution may be explicable by the nest structures and sources of incubation heat of each taxon. Although both megaloolithid and faveoloolithid eggs are found at mid-latitudes (Table 4.6), they might have been incubated with different sources of heat. The limited paleolatitudinal distribution of faveoloolithid eggs/eggshells (26.1°–47.1° from the equator) is consistent with the use of inorganic heat sources for their sandy in-filled hole nests. Sandy nests have a high thermal conductivity (Chabreck, 1973; Charruau et al., 2010; Charruau, 2012) but their incubation heat source would be either solar radiation or geothermal heat. In order to obtain enough heat for incubation, especially when using solar radiation, their distribution is limited to warm climates (e.g., warm temperate and arid climatic zones) at correspondingly lower latitudes. In contrast, sauropods laying megaloolithids probably used the organic origin of incubation heat in the mound nests, which potentially allowed them to nest at a higher latitude than sauropods laying faveoloolithids, although the distribution of megaloolithids is also currently known from mid-paleolatitudes (18.3°–44.9° from the equator). The latitudinal distributions of megaloolithids and faveoloolithids (up to 47° from the equator) are wider than those of living crocodylians and megapodes (up to 30° and 37° of the equator, respectively: Jones et al., 1995; Lance, 2003). The wider distributions of sauropod nesting sites could be related to the fact that Cretaceous was generally warmer than today, especially at higher latitudes (e.g., Huber et al., 2002); for example, the mean global surface temperature in Maastrichtian was estimated to 20–28 °C, which is 7–15°C higher than today (Upchurch et al., 2015). At the Auca Mahuevo nesting site in Patagonia (42.3°S of paleolatitude), where megaloolithid clutches are found, an average ambient temperature of a warm month was estimated to 40°C (Eagle et al., 2015), in which covered nests would have helped avoid overheating of eggs. Also, Campanian–Maastrichtian faveoloolithid eggs are found only a few degrees south from the Auca Mahuevo site (44.1°–44.9°S of paleolatitude) (Salgado et al., 2007). The warmer climate in Cretaceous thus would have allowed sauropods to maintain optimal incubation temperatures at a broader geographic range. Hadrosaur (Spheroolithidae) and maniraptoran theropod (Elongatoolithidae and Prismatoolithidae) eggs/eggshells also occur at a broad geographic/latitudinal ranges, although their incubation methods were likely different from one another (i.e., mound nests with microbial

87

respiration for hadrosaurs and open nests with brooding for maniraptorans). Hadrosaur and troodontid eggshells occur at extremely high paleolatitude (76.7°N) (Figure 4.11), where mean annual and summer temperatures were around 10 °C and 19 °C, respectively (Golovneva, 2000; Herman et al., 2016). This is equivalent to the climate in Frankfurt today (data derived from http://www.dwd.de/EN/Home/home_node.html), where successful nesting of a captive mound nesting megapode has been documented (Zoo Frankfurt: Baltin, 1969), supporting the idea that dinosaurs could also successfully nest using mounds in such conditions. Thus, mound nests, with their incubation heat generated by decaying plant/soil, would likely have been the only adequate covered nesting style, pending the availability of geothermal heat, for hadrosaurs at high paleolatitudes. Although the Kakanaut Formation is so far the only eggshell site at over 70° paleolatitude, footprints of young juvenile hadrosaurs (Fiorillo et al., 2014) are known from Cretaceous Alaska (72.4° paleolatitude according to Paleobiology Database), which may indicate that hadrosaurs nested at an even broader range in high arctic regions. Dinosaurs inferred to have open nests, such as maniraptorans, may have been brooders thus allowing them to nest in cooler environments as they would not have been dependent on nest materials and external sources for incubation heat. In summary, only dinosaur taxa capable of producing higher incubation heat, through plant decomposition in mounds, potentially geothermal activity, and adult contact incubation, were able to nest in high latitude environments.

88

Table 4.1. List of clutch specimens observed at the Royal Tyrrell Museum of Palaeontology (TMP), Drumheller, Alberta, Canada. Taxon/ootaxon Catalogue number Prismatoolithus levis (Troodon formosus) TMP 1994.179.01 P. levis (T. formosus) TMP 1996.086.01 P. sp. TMP 2008.075.51 Hypacrosaurus stebingeri TMP 1989.079.53 H. stebingeri TMP 1988.079.36 H. stebingeri TMP 1997.063.01

89

Table 4.2. List of ootaxa and their possible parental taxa used in this study. Oofamily Parental taxa Identification Reference based on: Dendroolithidae Torvosaurus and possibly Embryos Manning et al. (1997); (also known as therizinosauroids Kundrat et al. (2008); Phaceloolithidae) Araujo et al. (2013); Ribeiro et al. (2014) Elongatoolithidae Oviraptorosauria Embryos, a Dong and Currie gravid adult, (1996); Clark et al., and adults (1999); Norell et al. atop clutches (2001); Sato et al. (2005); Cheng et al. (2008); Weishampel et al. (2008); Fanti et al. (2012); Wang et al. (2016) Faveoloolithidae ?Sauropodomorpha Eggshell Grellet-Tinner and morphology Fiorelli (2010); Grellet-Tinner et al. (2012) Megaloolithidae Titanosaurs Embryos Chiappe et al. (1998, 2001) Prismatoolithidae Troodontids and Embryos and Varricchio et al. (1997, (excluding non-oviraptorosaur eggshell 2002); Zelenitsky and Preprismatoolithus) maniraptorans morphology Therrien (2008a); Lopez-Martinez and Vicens (2012) Spheroolithidae Ornithschian, including Embryos Horner and Makela Maiasaura and (1979); Hirsch and Hypacrosaurus Quinn (1990); Horner (1999, 2000)

90

Table 4.3. Results of one-way χ2-tests for the in-situ dataset of dinosaur ootaxa. Spheroolithidae was not tested due to the small sample size. Abbreviations: d.f., degree of freedom; n, sample size; NA, not applicable; p, probability for χ2-tests. Ootaxa n for n for Total n Expected n for Expected n for χ2 value d.f. p fine-grained coarse- fine-grained coarse-grained grained Dendroolithidae 5 5 10 5.288 4.712 0.033 1 0.855 Elongatoolithidae 14 13 27 14.277 12.723 0.011 1 0.915 Faveoloolithidae 4 32 36 19.037 16.963 25.207 1 <0.001 Megaloolithidae 43 22 65 34.372 30.628 4.596 1 0.032 Prismatoolithidae 9 11 20 10.576 9.424 0.498 1 0.480 Spheroolithidae 11 3 14 NA NA NA NA NA Other ootaxa 15 4 19 10.047 8.953 5.182 1 0.023 Total 101 90 191 NA NA NA NA NA

91

Table 4.4. Results of one-way χ2-tests for the ex-situ dataset of dinosaur ootaxa. Dendroolithidae was not tested due to the small sample size. Abbreviations: d.f., degree of freedom; n, sample size; NA, not applicable; p, probability for χ2-tests. Ootaxa n for n for Total n Expected n for Expected n for χ2 value d.f. p fine-grained coarse-grained fine-grained coarse-grained Dendroolithidae 1 3 4 NA NA NA NA NA Elongatoolithidae 15 15 30 16.599 13.401 0.345 1 0.557 Faveoloolithidae 7 9 16 8.853 7.147 0.842 1 0.359 Megaloolithidae 23 29 52 28.772 23.228 2.508 1 0.113 Prismatoolithidae 8 9 17 9.406 7.594 0.470 1 0.493 Spheroolithidae 13 8 21 11.619 9.381 0.387 1 0.534 Other ootaxa 42 15 57 31.538 25.462 7.769 1 0.005 Total 109 88 197 NA NA NA NA NA

92

Table 4.5. Number of occurrences of major dinosaur clutches among countries. Parentheses indicate number of pedogenic deposits in each case. Ootaxa Country Fine-grained deposits Coarse-grained deposits Elongatoolithidae China 10 (10) 2 (0) Mongolia 1 (0) 8 (1) Other countries 3 (1) 3 (2) Faveoloolithidae Argentina 0 (0) 8 (7) South Korea 3 (1) 21 (0) Other countries 1 (0) 3 (0) Megaloolithidae Argentina 5 (2) 5 (3) France 5 (0) 0 (0) India 0 (0) 10 (10) Romania 11 (10) 1 (1) Spain 22 (15) 6 (4) Prismatoolithidae Mongolia 1 (0) 8 (1) Canada and 6 (5) 1 (0) USA Other countries 2 (0) 2 (0)

93

Table 4.6. Ranges of absolute paleolatitudes in major dinosaur ootaxa. Abbreviation: n, sample size. Ootaxa n Range of absolute paleolatitudes Dendroolithidae 18 26.70° to 48.00° Elongatoolithidae 54 19.85° to 49.00° Faveoloolithidae 13 26.10° to 47.10° Megaloolithidae 28 18.30° to 44.85° Prismatoolithidae 33 19.95° to 76.70° Spheroolithidae 45 22.45° to 76.70°

94

Figure 4.1. Schematic diagram of tests and comparisons conducted in this study.

95

Figure 4.2. Boxplot of mean nest (Tnest) and ambient air temperature (Tair) among covered nests that use heat from microbial respiration, geothermal energy, and solar radiation. The range of nest temperature is indicated with the orange shadow.

96

Figure 4.3. Comparisons of nest materials/substrates between extant covered nests that use inorganic heat sources and organic (and some inorganic) heat sources: (a) barplots and (b) pie charts. Note that use of nest materials/substrates is different between these two types of incubation heat.

97

Figure 4.4. Comparisons of nest materials/substrates between extant mound and in-filled hole nests: (a) barplots and (b) pie charts. Note that use of nest materials/substrates is different between these two nest structures.

98

Figure 4.5. Comparisons of nest lithology among dinosaur oofamilies: (a) in-situ dataset and (b) ex-situ dataset. Note that in-situ nests of hadrosaurs and sauropods tend to be associated with particular type of sediments, whereas the other oofamilies are unrelated with certain lithology; oofamilies in the ex-situ dataset are not significantly related to certain lithology. Abbreviations: sig., significant; NS, not significant at 0.05 level.

99

Figure 4.6. Ratio of paleosol or pedogenetic features found in dinosaur nests. Note that nests of hadrosaurs and sauropods laying megaloolithids are strongly associated with paleosols, whereas other groups are not.

100

Figure 4.7. Boxplot of absolute paleolatitude in dinosaur egg/eggshell localities. Note that most localities range within 20° and 60° whereas spheroolithid and prismatoolithid eggshells are also known from further North (76.7°N).

101

Figure 4.8. Distributions of current covered nesters (i.e., crocodylians and megapodes): (a) mound nesters that use heat from microbial respiration, (b) in-filled hole nesters that use heat from geothermal energy, and (c) in-filled hole nesters that use heat from solar radiation.

102

Figure 4.9. Inferred nest structures and incubation heat sources for dinosaurs. Covered and open nest types inferred from eggshell porosity and/or taphonomy are from the literature (Tanaka et al., 2015 and references therein).

103

Figure 4.10. (a) Jurassic and (b) localities of dinosaur egg/eggshell remains. Paleomaps were modified from: http://cpgeosystems.com/index.html.

104

Figure 4.11. Late Cretaceous localities (dots) of dinosaur egg/eggshell remains: (a) hadrosaurs (Spheroolithidae) and sauropods laying megaloolithids, (b) sauropods laying faveoloolithids, and (c) Dendroolithidae, oviraptorosaurs (Elongatoolithidae) and troodontids (Prismatoolithidae). Paleomaps were modified from: http://cpgeosystems.com/index.html.

105

CHAPTER 5: ESTIMATION OF INCUBATION PERIOD IN DINOSAURS

5.1 INTRODUCTION Incubation period, measured as the time interval from oviposition to hatching, is important because it is related to various aspects of the life-history of archosaurs (e.g., altricial–precocial spectrum, embryonic growth, hatching synchrony, fledging period, adult and juvenile mortality rate: Heinroth, 1922; Lack, 1968; Drent, 1975; Ricklefs, 1993; Ricklefs and Starck, 1998; Martin, 2002; Deeming et al., 2006). Incubation period varies widely among living archosaur species, from about 10 to 100 days on average (Rahn and Ar, 1974; Birchard and Marcellini, 1996). Shorter incubation periods are considered advantageous in an evolutionary sense (Lack, 1968; Bosque and Bosque, 1995) because risk of nesting failure is expected to decrease with a shorter incubation period (Mayfield, 1961, 1975; Packard and Packard, 1994). Birds tend to have shorter incubation periods (10–90 days: Rahn and Ar, 1974) than those of crocodylians (68–99 days: Birchard and Marcellini, 1996), although it is unknown when this reduction occurred during archosaur evolution due to distant phylogenetic relationship between these groups. Non-avian dinosaurs are important to understand this shift in timing, although their incubation period requires estimation as it cannot be measured directly. Estimating incubation period in dinosaurs should be based on regression analyses of traits related to incubation period of extant archosaur species. A prior estimation of incubation period for dinosaurs (viz., sauropodomorphs) used only egg mass in simple linear regressions of crocodylians and birds (Ruxton et al., 2014). In crocodylians, there is no correlation between incubation period and egg mass (Deeming et al., 2006), whereas birds show a high correlation (r2 ≈ 0.7) (Rahn and Ar, 1974; Ar and Rahn, 1978; Deeming et al., 2006). Egg mass in crocodylians is thus not useful for predicting incubation period from a simple linear regression. Although there is a correlation in birds, the intercepts of regressions differ considerably among orders and thus prediction may not be accurate from a regression of all birds (Deeming et al., 2006). Furthermore, egg mass is not considered the single largest factor for influencing incubation period (Ricklefs and Stark, 1998; Martin, 2002; Deeming et al., 2006), as many other factors are known to affect their incubation period in birds and (e.g., incubation temperature, water vapor conductance of eggshell, embryonic growth rate, and phylogeny: Deeming et al., 2006). Considering the lack of

106

linear correlation in crocodylians and the multiple factors that influence incubation period, a method other than simple linear regression is more appropriate to estimate incubation period across Archosauria and within dinosaurs. Here I develop a model to predict incubation period across Archosauria using multiple regression analysis. Because this model will be applied to extinct taxa, the variables used are only those that influence incubation period and that can also be derived from the fossil record (i.e., egg mass, geometry of eggshell pore canals, and nest type). Using these variables, phylogenetic and non-phylogenetic (i.e., conventional) multiple regression analyses were applied to a large dataset of living bird and crocodylian species (n = 132) to obtain regression models for incubation period, which were then used to predict incubation period in dinosaurs. These predictions provide insight into how incubation period changed through the evolution of archosaurs.

5.2 MATERIALS AND METHODS In this study, reproductive parameters known to be related to incubation in extant archosaurs (i.e., egg mass, eggshell porosity, pore density, and nest type; Rahn and Ar, 1974; Hoyt, 1980; Deeming et al., 2006; Zimmermann and Hipfner, 2007) were investigated. Phylogenetic and conventional regression analyses were conducted to document the relationship between the aforementioned parameters and incubation period. Initially, simple bivariate linear regressions using egg mass were conducted in order to evaluate the validity of previous approaches (e.g., Ruxton et al., 2014), and then multiple regression models were conducted by adding more independent variables (i.e., eggshell porosity, pore density, and nest type). Subsequently, the same reproductive parameters were compiled for a number of eggs of non-avian dinosaurs, and incubation period for these taxa was estimated following the regression models.

5.2.1 Selection of living taxa

Incubation period (Ip, in day), egg mass (M, in g), eggshell porosity (i.e., total pore area in −1 −2 an egg divided by pore length, Ap∙Ls , in mm), pore density (D, in mm ), and nest type (NT) were compiled from 132 species of extant birds and crocodylians. As for previous studies (e.g., Tanaka and Zelenitsky, 2014a; Tanaka et al., 2015), eggshell of species with pore canals that approximate simple (non-branching) or tubular structures were included; species with irregular eggshell pore canals (e.g., Casuarius, Dromaius, Pterocnemia, Rhea, and Struthio: Tyler and Simkiss, 1959;

107

Board and Tullett, 1975; Board et al., 1977; Tyler and Fowler, 1979) were excluded from the −1 dataset because eggshell porosity could not be accurately calculated. Values of Ip, M, Ap∙Ls , D, and NT for extant species were taken from either the literature or from new measurements of egg specimens (Appendices 5.2–5.4).

5.2.2 Selection of fossil taxa/ootaxa −1 Data of M, Ap∙Ls , D, and NT were compiled for 23 extinct taxa and ootaxa (i.e., egg taxa), including titanosaurs, Lourinhanosaurus, oviraptorosaurs, and Troodon (Appendix 5.5). Selection criteria for the fossil taxa/ootaxa (e.g., eggshells with simple pore canals) followed Tanaka et al. (2015). Ootaxa examined in this study include Cairanoolithidae (possibly ornithischian), Megaloolithidae (sauropods), Elongatoolithidae (Oviraptorosauria), Prismatoolithidae (non-oviraptorosaur maniraptorans), and Continuoolithus (non-avian theropod) (Appendix 5.5). The ootaxon Protoceratopsidovum was not included because values provided in the literature are not consistent among studies (i.e., Sabath, 1991; Mikhailov et al., 1994).

5.2.3 Values for incubation period

Values of Ip for living archosaur species were compiled from the literature. Although Ip is defined as "the interval between the laying of the last egg of a clutch and the hatching of that egg" (Gill, 2007), it is difficult to obtain such values for each species because many studies do not report

Ip by strict definition. In this study, Ip values were averaged from any eggs within/among clutches (from wild individuals) for a single species, following the method by Deeming et al. (2006).

According to the compiled dataset in this study, Ip for brooding birds varies less than for species that build covered nests. For brooding birds, differences between mean Ip and minimal Ip values and between mean Ip and maximal Ip values, relative to the mean Ip values [i.e., 100·(mean

Ip – minimal Ip)/ mean Ip and 100·(mean Ip – maximal Ip)/ mean Ip, respectively], were, on average, 8.8% and 12.7%, respectively (n = 118, calculated based on Appendix 5.3). In contrast, incubation period of crocodylians and megapodes, which build covered nests, varies greatly. For example, reported Ip values range from 61 to 100 days in the Morelet's Crocodile (Crocodylus moreletii) (Platt et al., 2008) and from 50 to 90 days in the Mallee (Leipoa ocellata) (Frith, 1959). The wide range of incubation period is most likely related to fluctuations in incubation temperate in these species (27–39 °C in megapodes: Booth and Jones, 2002, and 23–41 °C in crocodylians:

108

Staton and Dixon, 1977; Ferguson, 1985), because their nest temperature depends on environmental conditions (i.e., solar radiation, geothermal activity, and microbial respiration) (Ferguson, 1985; Booth and Jones, 2002). Prior research on artificially incubated eggs of covered nesters has revealed that there is a negative relationship between Ip and incubation temperature (e.g., Webb et al., 1987; Lang et al., 1989; Lang and Andrews, 1994; Cornejo et al., 2014). Normal embryonic development and the highest hatchability are shown to be achieved at 28–34 °C in crocodylians ( and

Seymour, 1990) and at 32–34 °C in megapodes (Booth, 1987; Cornejo et al., 2014). Thus, Ip values for covered nesters were taken from eggs that were artificially incubated at these ideal temperature ranges, similar to the method by Birchard and Marcellini (1996). Crocodylian species for which Ip values from artificially incubated eggs were not available, Ip values from naturally incubated eggs (either from wild or captive ) at these temperature ranges were used. It was confirmed that there is an agreement of Ip values between artificially incubated eggs at these temperature ranges and naturally incubated eggs in which mean incubation temperature falls within the ranges

(Appendices 5.1–5.4), and there is no significant difference of mean Ip values between them (paired-samples T-test: p = 0.592: Appendix 5.6).

5.2.4 Selection of independent variables Egg mass, eggshell porosity, pore density, and nest type were selected as independent variables for regression analyses because these parameters are directly or indirectly related to incubation period in extant archosaurs (Hoyt, 1980; Deeming et al., 2006; Zimmermann and

Hipfner, 2007). According to Carey (1983), Ip can be expressed as:

[5-1] where F is the fraction of water vapor lost during incubation, relative to M, and MH2O is the daily −1 loss of water vapor (mg H2O∙day ). Mean values of F range from 0.10 to 0.16 in birds that build covered nests (i.e., crocodylians and megapodes: Lutz and Dunbar-Cooper, 1984; Seymour et al., 1987), values of which are consistent and within the range of those for open nesters (0.10 to 0.23 with the mean of 0.15: Ar and Rahn, 1980). Thus, the difference of F between covered and open nesters was assumed insignificant and can be considered as a constant in archosaurs. The daily loss of water vapor can be expressed as:

[5-2]

109

where Gex is the experimentally-derived water vapor conductance (i.e., diffusive capacity) of an −1 −1 egg (mg H2O∙day ∙Torr ) and ∆P is the gradient of water vapor pressure between the inside of the egg and the nest (Torr) (Ar et al., 1974). Thus, Ip is expressed as:

[5-3]

Additionally, using the pore canal geometry, water vapor conductance can be deduced from Fick's first law of gas diffusion:

[5-4]

−1 −1 where c is a unit conversion constant (mg H2O∙sec∙day ∙mol ), DH2O is the diffusion coefficient 2 −1 of water vapor (mm ∙sec ), Gmo is the morphometrically-derived water vapor conductance (mg −1 −1 3 −1 −1 H2O∙day ∙Torr ), R is the universal gas constant (mm ∙Torr∙mol ∙K ), and T is the absolute temperature of incubation (°K) (Ar et al., 1974). Because DH2O, T, c, and R are considered as constant (Tanaka and Zelenitsky, 2014a), Equation 5-4 can be simplified as:

[5-5]

According to Tanaka and Zelenitsky (2014a), Gex and Gmo in logarithmic scale show a significant positive correlation:

[5-6] −1 As a consequence, Equations 5-3, 5-5 and 5-6 explain Ip is related to M, Ap∙Ls , and ∆P. Variables −1 of M and Ap∙Ls are thus included in the regression models, although ∆P is not estimated from the fossil record. As a proxy for ∆P, nest type (categorical variable: covered nest vs. open nest) was included in the regression models because values of ∆P should be low in covered nests and high in open nests (Tanaka et al., 2015), suggesting that nest type (i.e., covered and open nest) can reflect ∆P. Pore density (D, in mm−2) was also included in the analyses in order to consider the pore aperture resistance, which also affects the rate of water vapor diffusion. According to Stefan's law of gas diffusion, resistance against diffusive water vapor increases when pore diameter is large relative to pore length (Rahn et al., 1978; Paganelli, 1991). This phenomenon alters the rate of water vapor diffusion as the diffusion rate becomes lower in relatively thick pores than in narrow −1 pores (Rahn et al., 1978). Thus, it can be hypothesized that even though Ap∙Ls and M are the same,

Ip is relatively longer in eggs consisting of fewer large pores than those containing many small pores. As a proxy of this effect, pore density was also included in the regression analyses.

110

5.2.5 Values for egg mass, eggshell porosity, and pore density −1 Values of M, Ap∙Ls , and D for both living and fossil taxa/ootaxa were measured and/or compiled following (Tanaka and Zelenitsky, 2014a; Tanaka et al., 2015). Eggshell specimens measured for this study are listed in Appendix 5.2. Estimates of egg mass for extinct taxa/ootaxa are based on egg length (mm) and breadth (mm) following Hoyt (1979). When values of D are not provided in the literature, the total number of pores in an egg was divided by eggshell surface area (mm2), which is derived from egg length (mm) and breadth (mm).

5.2.6 Nest type Nest type (NT) was compiled from the literature for living archosaur species and was inferred based on Tanaka et al. (2015) (Chapter 3) for extinct archosaur taxa/ootaxa. Nest type was classified into either covered nests, where the eggs are completely covered with nest materials/substrates, or open nests, where the eggs are partially or fully exposed, following Tanaka and Zelenitsky (2014b). Nest data for living species were obtained from Tanaka et al. (2015) with additional papers (Harrison, 1975; Lomholt, 1976; Davis et al., 1984; del Hoyo et al., 1992; Baicich and Harrison, 1997). Nest type of extinct taxa/ootaxa was based on Tanaka et al. (2015), which inferred dinosaur nest types using estimates of egg mass and eggshell porosity. In this study, Cairanoolithidae, Megaloolithidae, Continuoolithus, and Lourinhanosaurus were classified into covered nest, whereas Elongatoolithidae, Prismatoolithus, and Sankofa were classified into open nest. Nest type was included as a categorical variable (i.e., covered nest coded as “0” and open nest coded as “1”).

5.2.7 Statistical analyses −1 Incubation period was correlated with M, Ap∙Ls , D, and NT for 132 living archosaur species, using regression analyses of non-phylogenetic ordinary least-squares (OLS) and phylogenetically-generalized least-squares (PGLS). The phylogenetic approach, which considers phylogenetic relationships of the species within the dataset, was conducted because the assumption of the independence of data points (i.e., species) may inflate the chance of type I errors (e.g., Garland et al., 2005). In order to compare these regression models with the previous approach based on egg mass (i.e., simple linear regression model by Ruxton et al., 2014), both

111

simple linear and multiple regressions were conducted. Simple linear regressions were applied to −1 scatterplots of Ip as a function of M, Ip as a function of Ap∙Ls , and Ip as a function of D. For multiple regression analyses, various combinations of independent variables were tested: 1) −1 eggshell porosity variables only (i.e., Ap∙Ls and D), 2) eggshell porosity variables and nest type −1 −1 (i.e., Ap∙Ls , D, and NT), 3) egg mass and eggshell porosity variables (i.e., M, Ap∙Ls , and D), and −1 4) all variables (i.e., M, Ap∙Ls , D, and NT). The regression analyses, procedures outlined by Campione and Evans (2012), were implemented with "gls()" function under the Brownian motion model (function "corBrownian") using software platform R3.1.3 (http://www.r-project.org/) and R packages "ape" and "nlme". Additional R packages, such as "car" and "MASSTIMATE", were used to calculate the variance inflation factor (VIF) and standard error of the estimate (SEE). Multicollinearity, which occurs due to intercorrelations among independent variables, was tested with VIF values; generally a VIF value greater than 10 indicates multicollinearity (Kutner et al., 2005). Prior to analyses, all values were transformed into 10-based logarithms as logarithmic distribution is common in nature (e.g., Zar, 1968; Sokal and Rohlf, 1995). Simple linear and multiple regression models were evaluated with SEE, the coefficient of determination (r2), the percent prediction error that is calculated from non-logarithmic values (PPE), and the Akaike Information Criterion (AIC) that is used for evaluation of model fits; lower AIC value indicates a better model (Lavin et al., 2008; Swanson and Garland, 2009; Gartner et al., 2010). A phylogenetic tree of 132 species of living archosaurs was reconstructed for phylogenetic regression analyses (Appendix 5.12), using the PDAP module v.1.16 (Midford et al., 2010) of the software Mesquite 3.02 (Maddison and Maddison, 2010). The phylogenetic tree of Tanaka et al. (2015) (Chapter 3) was modified, following phylogenies of Fjeldsa (2004), Thomas et al. (2004), Gonzalez et al. (2009), and Jarvis et al. (2014). Following Tanaka et al. (2015), branch length was approximated using divergence times of each node. Because branch length was not adequately standardized with divergence time, an additional arbitrary standardized method, such as the transformation of branch lengths by Nee's method, was also used, following Garland et al. (1992). Subsequently, multiple regression models obtained from living species were applied for extinct taxa/ootaxa and their incubation period was estimated. The range of estimations for each taxon/ootaxon was calculated from mean PPE values. Estimates of Ip and M in dinosaurs were compared on a bivariate plot with data points of living archosaur species. Values of Ip in sauropods

112

(i.e., Megaloolithidae) and non-avian maniraptorans (i.e., Elongatoolithidae, Prismatoolithus, and Sankofa) were regressed with egg mass when these variables were significantly correlated (p < 0.05). Ordinary least-squares regression was used but phylogenetic regressions were not applied for these taxa/ootaxa due to the uncertainty of phylogenetic positions of these ootaxa. Regressions of living bird groups (i.e., birds overall and ), taken from Deeming et al. (2006), were also included in bivariate plots.

5.3 RESULTS Simple linear regression analyses revealed that egg mass is not an accurate predictor of incubation period in living archosaurs (r2 = 0.28: Table 5.1; Figure 5.1a). Values of AIC were lower in PGLS models (-337.02 to -257.39) than in the OLS model (-82.80), suggesting that the phylogenetic regressions are better fits. Values of mean PPE vary from 33.75 to 40.45%, indicating that estimated Ip values predicted by the regressions of Ip against M differ, on average, by 34 to

40% relative to the actual Ip values, which is probably due to the variations of Ip among bird/crocodilian families. −1 −1 Simple linear regressions of Ip against Ap∙Ls and Ip against D revealed that Ap∙Ls is 2 positively correlated with Ip, while D is negatively correlated (r > 0.4 in both cases: Table 5.1;

Figures 5.1b and 5.1c). Like the results of the regressions of Ip against M, PGLS models were better than OLS models in terms of the lower AIC values. Although the lowest AIC value was 2 found in the regression of Ip against M (PGLS1 of Table 5.1), values of r were higher and values −1 of PPE were generally lower in the regressions of Ip against Ap∙Ls and Ip against D than those of Ip against M. Based on the results of simple linear regressions, combinations of independent variables −1 in multiple regressions were determined. The first combination (i.e., Ap∙Ls and D) includes only variables that are related to the pore canal geometry. This is the simplest combination but the two continuous variables show linear correlations with Ip, much stronger than the correlations of M with Ip. The second combination includes these two continuous variables and nest type (NT) as a categorical variable because NT can be a proxy for the gradient of water vapor pressure between the inside of the egg and the nest (see "5.2.4 Selection of independent variables"). In the third −1 combination, all continuous variables were included (i.e., M, Ap∙Ls , and D). Although M is poorly correlated with Ip, M theoretically affects Ip as well (see "5.2.4 Selection of independent variables").

113

The last combination includes all continuous variables and nest type (NT).

Multiple regression analyses revealed that Ip is moderately to highly correlated to −1 incubation parameters (i.e., M, Ap∙Ls , D, and NT) in living archosaurs (r = 0.78–0.91: Table 5.2). −1 The correlations of each variable were generally significant (p < 0.05), but Ap∙Ls and NT were not significant in seven models (Table 5.2). Multicollinearity was assumed not to be significant because VIF values were less than 5.0 in all models. Mean PPE values were less than 20% in all except for the two models (28.67% and 33.43% for PGLS3 and PGLS4, respectively), which are considerably lower than those of most simple linear regression models (Figure 5.2). As a general trend, AIC values decreased with the increase of the number of independent variables (Table 5.2). Additionally, OLS models tend to have higher r2, lower SEE, and lower mean PPE values than PGLS models. Between PGLS models, PGLS adjusted by Nee's method was better than PGLS adjusted with divergent time in terms of higher r2, lower SEE, and lower mean PPE values. However, lower AIC values were generally obtained in the opposite order: PGLS adjusted with divergence time, PGLS adjusted by Nee's method, and OLS. The lowest AIC value was found in the PGLS model adjusted with the divergence time (-337.90: PGLS9 of Table 5.2) that includes all −1 independent variables (i.e., M, Ap∙Ls , D, and NT).

Based on these multiple regression models, Ip of extinct archosaurs was estimated for selected taxa (Table 5.3) and for all ootaxa (Appendices 5.7 and 5.8). Estimated Ip values ranged from 42 to 120 days for possible ornithischians (i.e., Cairanoolithidae), from 36 to 136 days for sauropods (i.e., Megaloolithidae), and from 32 to 93 days for non-avian maniraptorans (i.e., Elongatoolithidae, Prismatoolithus, and Sankofa). According to the aforementioned best fit model

(i.e., PGLS9), Ip of titanosaurs, Lourinhanosaurus, oviraptorosaurs, and Troodon were estimated to 102, 102, 60, and 49 days, respectively (Figure 5.3). Estimates of Ip in extinct taxa varied widely depending on which regression model is used, yet distributions of the data points in the bivariate plots of log Ip against log M generally are within the range of living birds (Appendices 5.13–5.23).

Most of the estimated Ip values, relative to egg mass, in extinct taxa are shorter than Ip of living crocodylians, although these values are comparable to Ip of many living bird species with equivalent egg mass. Among extinct taxa, Lourinhanosaurus had a relatively long Ip, generally above or near the regression of Procellariiformes, which shows among the longest Ip in birds

(Boersma, 1982; Grant et al., 1982c; Rahn and Whittow, 1988; Deeming et al., 2006). Estimated Ip values of sauropods (Megaloolithidae) and a possible ornithischian (Cairanoolithidae) varied from

114

below the regression of birds overall and up to the regression of Procellariifomes, although estimates from two models (i.e., PGLS10 and PGLS14) were plotted just above the regression of Procellariiformes. The OLS regressions of non-avian maniraptorans (i.e., Elongatoolithidae, Prismatoolithus, and Sankofa) were generally just below the regression of Procellariiformes but above that of birds overall.

5.4 DISCUSSION This study reveals that incubation period across Archosauria can be more accurately estimated with multiple parameters of incubation traits (i.e., egg mass, pore canal geometry, and nest type) than with a single trait such as egg mass, eggshell porosity, or pore density. As a general trend, multiple regression models in this study showed higher r2 values and lower mean PPE values than the simple linear regression models. Likewise, values of AIC and SEE tend to be lower in the multiple regression models than the simple linear regression models. For example, the predictive power of our best multiple regression model (i.e., the model of the lowest AIC: PPE of 17%) was over twice as high as that of the simple linear regression models (i.e., PPE of 30 to 40%, except for two cases that are 21 and 24%). These results indicate that multiple regression methods are useful to estimate incubation period of archosaurs, including extinct taxa. Estimates of incubation period for extinct archosaurs were consistent among the nine multiple regression models conducted in this study. The bivariate plots reveal that the trend of the distribution of extinct taxa/ootaxa is relatively consistent, although estimates of incubation period in each extinct taxon/ootaxon varies among the models (Figures 5.3; Appendices 5.13–5.23). As a general trend, estimates of incubation period, relative to egg mass, for most non-avian dinosaurs were shorter than that of crocodylians but were comparable to that of extant bird species. These results differ from the interpretation of Ruxton et al. (2014), based on regressions of incubation period and egg mass in crocodylians and birds, where incubation of sauropods was long. The relatively short incubation period of most dinosaur groups analyzed could be the result of relatively low eggshell porosity, high pore density, and/or open nest type (Figure 5.4). Results of multiple regression analyses indicate that egg mass and eggshell porosity are positively correlated and pore density is negatively correlated with incubation period. Also, incubation period is estimated to be shorter in open nest types where the gradient of water vapor pressure between the inside of the egg and its nest is high, and longer in covered nest types where the

115

gradient is low. In living archosaur species, crocodylians are covered nesters that have a high eggshell porosity and a low pore density, which contribute to a long incubation period. Most birds, on the other hand, have shorter incubation periods due to a lower eggshell porosity, a higher pore density, and an open nest type. A relatively short incubation period of sauropods and possibly some ornithischians shown in most regression models could be due to a high pore density even though eggshell porosity is high and the nest type is covered (Tanaka et al., 2015). A similar adaptation is found in living megapodes, grebes and , which share such pore system features (i.e., relatively high pore density and high eggshell porosity: Tullett and Board, 1977; Davis et al., 1984; Ar and Rahn, 1985; Tanaka et al., 2015) with these dinosaurs. Although grebes and loons build open nests in aquatic environments, their nests are humid, like the nests of megapodes and crocodylians (Davis et al., 1984; McIntyre et al., 1997). Relatively short incubation period of non-avian maniraptorans is comparable to that of many bird species probably because they share a low eggshell porosity, a high pore density, and an open nest type. Lourinhanosaurus is shown to have had a relatively longer incubation period than other dinosaurs in this study because its nest and eggshell traits (i.e., high eggshell porosity, low pore density, and covered nest type) are comparable to those of living crocodylians. These results suggest that during archosaur evolution, the major reduction of incubation period should have occurred in clades between Crurotarsi and Dinosauria, where there is a large gap in our knowledge of the incubation period of these (Figure 5.5). Mean incubation period of Crocodylia is 2.5 times longer than that of birds with equivalent egg mass; however, incubation period of possible ornithischians and sauropods is significantly shorter and comparable to that of living birds. Among Dinosauria, including Aves, there may have been several independent increases and decreases of the relative length of incubation. For example, comparison of incubation period in living birds overall shows that estimates of possible ornithischian and sauropod clades are slightly lower, but that of Lourinhanosaurus is higher. Within Aves, incubation period independently became relatively longer in multiple taxa (e.g., kiwis, megapodes, and pelagic : Calder, 1979; Seymour and Ackerman, 1980; Boersma, 1982; Grant et al., 1982c; Rahn and Whittow, 1988) and shorter in others (e.g., altricial birds: Deeming et al., 2006). Assuming a precocial developmental mode in dinosaurs (e.g., Geist and Jones, 1996; Norell et al., 2001; Varricchio et al., 2002; Kundrat et al., 2008; Weishampel et al., 2008; Rogers et al., 2016), incubation period can be linked to growth rate of the embryos. Research on embryonic

116

osteology in dinosaurs suggests that the hatchlings were likely precocial, indicating that they were more developed than altricial hatchlings (e.g., Geist and Jones, 1996; Norell et al., 2001; Varricchio et al., 2002; Kundrat et al., 2008; Weishampel et al., 2008; Rogers et al., 2016; but contra. to Horner et al., 2001; Reisz et al., 2005). Embryonic growth rates of dinosaurs are generally inferred to be high due to highly vascularized embryonic bone tissues (e.g., hadrosaurs, sauropodomorphs, Lourinhanosaurus, oviraptorosaurs, and Troodon: Castanet et al., 2000; Horner et al., 2001; de Ricqlès et al., 2001; Weishampel et al., 2008; Garcia and Cerda, 2010; Reisz et al., 2013; Wang et al., 2016) (Figure 5.5). Thus, a relatively short incubation period revealed here in most taxa, in addition to highly vascularized embryonic bone shown in other studies, suggests that dinosaur embryos grew at a relatively fast rate. Additionally, relatively short incubation period along with a large clutch size in dinosaurs could have been selected for to offset high nest predation pressure. Among living archosaur species, one of the major factors of nesting failures is depredation of nests (e.g., Skutch, 1985; Mazzotti, 1989; Martin, 1993). Nest predation was also probably common in dinosaurs although direct evidence of this is scarce (Wilson et al., 2010; Venczel et al., 2015). In order to reduce the chance of nest predation, shorter incubation is favorable (Lack, 1968; Bosque and Bosque, 1995) and a large clutch size can increase the odds of nesting success (i.e., percentage of nests that produce at least one hatchling from the total number of nests: Mazzotti, 1989). Clutch size of dinosaur groups was fairly large: e.g., 16–20 for hadrosaurs (Horner, 1999), 25–40 for sauropods (Vila et al., 2010b), 34 for oviraptorosaurs (KT pers. obs.), and 24 for troodontids (Varricchio et al., 1997). These values are comparable to that of crocodylians (12–55: Ferguson, 1985; Thorbjarnarson, 1996) but larger than those of birds (1–20: Gill, 2007) (Figure 5.5). Furthermore, it has been suggested that hadrosaurs and sauropods had multiple clutches per (Werner and Griebeler, 2013). Thus, dinosaurs possibly retained a large clutch size as in crocodylians but significantly reduced incubation period as birds, both of which may have helped reduce the chance of nesting failure. These features suggest that non-avian dinosaurs had a life history pattern with relatively small eggs, large number of offspring, and rapid embryonic development. Eggs of non-avian dinosaurs were very small compared with their body masses (e.g., 0.040–0.142% of body mass in ornithischians, 0.008–0.073% in sauropods, and 0.342–0.748% in non-avian theropods), values of which are comparable to or even smaller than those in crocodylians (mean 0.31%) and much

117

smaller than those in birds (mean 6.55%) (Figure 5.5: values calculated from Tables S1 and S2 of Werner and Griebeler, 2013), although relative egg mass increased in the course of theropod evolution. Instead, clutch size of non-avian dinosaurs was generally large (20–40) and some dinosaur clades (e.g., hadrosaurs and sauropods) may have had more than one clutch in a year (Werner and Griebeler, 2013). Furthermore, hatchlings were likely precocial regardless of relatively short incubation. These reproductive features could be beneficial to increase reproductive rate and fitness in an environment where offspring mortality rate is high.

118

Table 5.1. Results of simple linear regression analyses in extant archosaur species. −1 Abbreviations: AIC, Akaike Information Criterion; Ap∙Ls , eggshell porosity; CI, confidence interval; D, pore density; M, egg mass; OLS, ordinary least-squares; PGLS, phylogenetically-generalized least-squares; PPE, percent prediction error; r2, coefficient of determination; SEE, standard error of the estimate. PGLS1 and PGLS3 are PGLS models with branch lengths determined by divergence time, and PGLS4 to PGLS6 are PGLS models adjusted with Nee's method. The sample sizes of all models are 132. Model Intercept Coefficient r2 AIC SEE PPE (%) −1 log M log Ap∙Ls log D Mean Lower Upper 95% CI 95% CI

OLS1 1.000 0.295 0.275 - 82.796 0.162 33.753 26.755 40.751 OLS2 1.361 0.243 0.416 -110.152 0.150 30.189 25.528 34.850 OLS3 1.471 -0.447 0.631 -170.481 0.119 21.045 16.746 25.344 PGLS1 1.322 0.196 0.275 -337.020 0.174 40.449 37.310 43.588 PGLS2 1.617 0.066 0.416 -281.868 0.214 39.781 36.693 42.870 PGLS3 1.637 -0.108 0.631 -276.688 0.220 37.687 34.866 40.509 PGLS4 1.028 0.297 0.275 -257.388 0.134 34.141 28.192 40.090 PGLS5 1.473 0.124 0.416 -190.620 0.172 33.791 29.039 38.543 PGLS6 1.515 -0.293 0.631 -201.409 0.166 24.453 20.608 28.297

119

Table 5.2. Results of multiple regression analyses in extant archosaur species. −1 Abbreviations: AIC, Akaike Information Criterion; Ap∙Ls , eggshell porosity; CI, confidence interval; D, pore density; M, egg mass; NT, nest type; OLS, ordinary least-squares; PGLS, phylogenetically-generalized least-squares; PPE, percent prediction error; r2, coefficient of determination; SEE, standard error of the estimate. PGLS7 to PGLS10 are PGLS models with branch lengths determined by divergence time, and PGLS11 to PGLS14 are PGLS models adjusted with Nee's method. The sample sizes of all models are 132. Note −1 2 that * and † indicate log Ap∙Ls and NT were not significant (p > 0.05), respectively. The highest r and the lowest AIC, SEE, and PPE are in bold. Model Intercept Coefficient r2 AIC SEE PPE (%) −1 log M log Ap∙Ls log D NT Mean Lower Upper 95% CI 95% CI

OLS4 1.369 0.165 - 0.369 0.802 - 243.123 0.088 16.063 13.321 18.805 OLS5† 1.435 0.149 -0.331 -0.059 0.805 -238.691 0.087 15.724 12.946 18.502 OLS6 1.233 0.093 0.119 -0.372 0.815 -244.858 0.085 15.554 12.989 18.120 OLS7* 1.318 0.159 0.043 -0.272 -0.161 0.834 -251.749 0.081 14.446 11.925 16.967 PGLS7 1.532 0.087 -0.162 0.799 -296.206 0.198 28.666 26.500 30.833 PGLS8 1.671 0.081 -0.160 -0.259 0.759 -297.957 0.194 16.556 12.857 20.255 PGLS9* 1.289 0.189 0.006 -0.104 0.612 -334.514 0.166 33.429 30.947 35.911 PGLS10* 1.424 0.187 0.000 -0.102 -0.248 0.719 -337.903 0.162 17.355 14.882 19.828 PGLS11 1.384 0.156 -0.351 0.802 -242.089 0.138 16.016 13.346 18.686 PGLS12† 1.516 0.144 -0.336 -0.155 0.800 -240.531 0.137 16.921 13.926 19.916 PGLS13* 1.088 0.222 0.044 -0.228 0.740 -272.425 0.120 19.800 16.883 22.716 PGLS14* 1.258 0.236 0.020 -0.199 -0.221 0.820 -276.749 0.116 14.951 12.250 17.651

120

Table 5.3. Estimates of incubation period ± prediction errors (days) in dinosaurs based on multiple regression models. Prediction errors were calculated based on mean percent prediction errors in Table 5.2. See Table 5.2 for abbreviations. Model Titanosaurs Lourinhanosaurus Oviraptorosauria Troodon formosus (Megaloolithus antunesi (Macroolithus (Prismatoolithus patagonicus) yaotunensis) levis) Estimate Error Estimate Error Estimate Error Estimate Error OLS4 66.832 ±10.735 95.953 ±15.413 74.456 ±11.960 56.939 ±9.146 OLS5 70.241 ±11.045 97.172 ±15.280 67.491 ±10.612 52.981 ±8.331 OLS6 71.751 ±11.160 100.205 ±15.586 84.269 ±13.107 62.469 ±9.717 OLS7 86.429 ±12.486 106.940 ±15.449 70.267 ±10.151 54.745 ±7.908 PGLS7 59.233 ±16.980 68.690 ±19.691 60.328 ±17.294 52.502 ±15.050 PGLS8 78.287 ±12.961 90.963 ±15.060 44.325 ±7.339 38.945 ±6.448 PGLS9 78.232 ±26.152 78.167 ±26.130 80.391 ±26.874 65.045 ±21.744 PGLS10 101.991 ±17.701 102.199 ±17.737 59.728 ±10.366 48.793 ±8.468 PGLS11 65.387 ±10.472 92.338 ±14.789 72.624 ±11.631 56.343 ±9.024 PGLS12 82.120 ±13.895 114.674 ±19.404 64.151 ±10.855 50.698 ±8.579 PGLS13 80.157 ±15.871 89.497 ±17.720 87.044 ±17.234 63.695 ±12.611 PGLS14 112.344 ±16.796 121.631 ±18.185 73.775 ±11.030 55.223 ±8.256

121

Figure 5.1. Bivariate plots between incubation period and predictor variables (i.e., egg mass, eggshell porosity, and pore density). Regression lines are based on the simple linear regressions in Table 5.1.

122

Figure 5.2. Comparison of percent prediction error (%) of incubation period estimates among regression models. Purple circles and solid bars represent mean and 95% confident intervals, respectively. See Tables 5.1 and 5.2 for details of the models.

123

Figure 5.3. Bivariate plot of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa. Estimates of incubation period in non-avian dinosaurs, which are derived from PGLS10 in Table 5.2, tend to be relatively shorter than incubation period of many crocodylian species. The regressions of birds overall (n = 1490) and Procellariiformes (n = 75) were drawn based on Deeming et al. (2006). The regression of Procellariiformes represents an example of a bird group that shows extremely long incubation period in birds. The regressions of Megaloolithidae (n = 12, r2 = 0.700) and extinct maniraptorans (n = 7, r2 = 0.944) conducted here based on ordinary least-squares are parallel to and close to the regression of Procellariiformes.

124

Figure 5.4. Comparisons of predictor variables in living and extinct archosaur taxa/ootaxa, including some outlier taxa of living bird groups.

125

Figure 5.5. Transition of incubation period in archosaurs. (a) Phylogenetic tree of archosaurs with possible transitions of reproductive traits. (b) Bar plot with error bars (standard error) of the relative incubation period, which is calculated from the actual (for living species) or estimated incubation period (extinct taxa) divided by predicted incubation period from the regression of birds overall in Deeming et al. (2006). Estimates of incubation period for extinct taxa were derived from PGLS10 in Table 5.2. (c) Bar plot with error bars (standard error) of clutch size in each taxon. Note that the major reduction of incubation period should have occurred in clades between (/Crurotarsi) and Dinosauria, whereas that of clutch size between non-avian theropods and Aves. (d) Bar plot with error bars (standard error) of relative egg mass, values of which are residuals of log egg mass from the bird regression of log egg mass and log adult body mass. Values of egg mass, adult body mass, and the regression were derived from Werner and Griebeler (2013). Note that the relative egg mass gradually increased in theropods.

126

CHAPTER 6: GENERAL DISCUSSION

The findings of this dissertation provide insight into the evolution of reproductive traits among archosaurs (Figure 6.1), with respect to nest type, incubation heat source, incubation behavior, and incubation period. New methods were developed in this dissertation, largely based on data related to eggs, eggshells, and nests of living archosaurs, to infer these reproductive traits in non-avian dinosaurs. The covered nest type, which is found in all extant crocodylians, was inferred for more basal dinosaur clades, including ornithischians, sauropodomorphs, and the theropod Lourinhanosaurus, based on analysis of egg and eggshell parameters (Chapters 3 and 4). Because they used covered nests, these dinosaur taxa likely incubated their eggs with external heat sources. Distribution and lithologies of clutches suggest that some sauropodomorphs (faveoloolithids) obtained incubation heat from solar radiation or geothermal activity, whereas other sauropodomorphs (megaloolithids) and some hadrosaurs generally used microbial respiration (Chapter 4). Derived theropods (e.g., oviraptorosaurs and troodontids) are inferred to have had a partially open nest type, based on eggshell porosity and taphonomic studies, and are thus inferred to have practiced brooding behavior (Chapter 3). Regardless of the nest type, incubation heat sources, and incubation behaviors, incubation periods of most non-avian dinosaur clades were relatively shorter than crocodylians, more comparable to the incubation period of birds (Chapter 5). The results of this dissertation are consistent overall with previous studies that have also shown that non-avian theropod dinosaurs acquired bird-like reproductive features and behaviors prior to the , for example, partially open nests (Clark et al., 1999; Varricchio et al, 2013), asymmetrical eggs (e.g., Hirsch and Quinn, 1990; Varricchio et al., 1997), pigmented eggs (Wiemann et al., 2015), three-layered eggshell (e.g., Varricchio and Jackson, 2004), and monoautochronic (i.e., sequential ovulation: Varricchio et al., 1997; Sato et al., 2005). The results of this dissertation, along with previous studies, suggest that reproductive style and hence nesting strategy differs among major archosaurian clades and that a transition to more bird-like nesting methods occurred through archosaur evolution. Crocodylians show a plesiomorphic style, characterized by covered nest type, long incubation, relatively small egg size, and large clutch size. Furthermore, their hatchlings are considered precocial (Whitehead and

127

Seymour, 1990). Reproductive style of relatively primitive dinosaurs (i.e., ornithischians, sauropodomorpha, and the primitive theropod Lourinhanosaurus) is similar to that of crocodylians, but sauropodomorphs and possibly ornithischians may have attained a relatively short incubation period, which could have contributed to a reduction in egg mortality and thus increased reproductive efficiency. Probably within non-avian maniraptorans, nest type changed from covered to partially open, which may have been associated with brooding behavior (Varricchio et al, 2013; Chapter 3), and thus the degree of parental care during incubation presumably increased in this clade. Then, a decrease in clutch size (Chapter 5), the loss of a functional (Varricchio et al., 1997; Sato et al., 2005; Zheng et al., 2013), the increase of relative egg size (Varricchio et al., 1997; Chapter 5), egg turning behavior during incubation (Varricchio et al., 1997), and altricial developmental mode (Starck and Ricklefs, 1998) occurred within Aves or close to the origin of Aves (Figure 6.1); thus, there was a shift from relatively small eggs with a large clutch size to relatively large eggs with a small clutch size. Therefore, birds, in general, tend to produce fewer offspring but probably provide more parental care than crocodylians and non-avian dinosaurs that tend to have larger clutches and some degree of parental care. Although this interpretation oversimplifies the diverse nesting strategies of archosaurs, it can be concluded that such research broadens our understanding of the evolution of reproductive strategies among archosaurs.

128

Figure 6.1. Transition of reproductive traits in archosaurs. Orange and green boxes indicate plesiomorphic and derived conditions, respectively. References: 1, Chapter 4 in this dissertation; 2, Chapter 3 in this dissertation; 3, Varricchio et al. (1997); 4, Sato et al. (2005); 5, Grellet-Tinner et al. (2006); 6, Zelenitsky and Therrien (2008a); 7, Wiemann et al. (2015); 8, Varricchio and Jackson (2004); 9, Jackson et al. (2010); 10, Xu et al. (2014); 11, Zheng et al. (2013); 12, Geist and Jones (1996); 13, Norell et al. (2001); 14, Varricchio et al. (2002); 15, Weishampel et al. (2008); 16, Rogers et al. (2016); 17, Horner (1982); 18, Mateus et al. (1998); Chiappe et al. (2001, 2004); 20, Cunha et al. (2004); 21, Reisz et al. (2012); 22, Varricchio et al. (2015). 129

CHAPTER 7: CONCLUSIONS

Knowledge about the types of nests, incubation behavior (e.g., brooding), incubation heat sources, and incubation period of non-avian dinosaurs can provide insight into the evolution of nesting and egg incubation among archosaurs. However, none of these traits (i.e., nest type, incubation behavior, incubation heat source, and incubation period) are directly observable or measurable from the fossil record because of the low or lack of preservation potential. Some information related to these behaviors thus can be gleaned indirectly through comparisons with extant archosaurs (i.e., crocodylians and birds). By comparing fossil eggs and eggshells with extant species, Chapters 2 to 5 investigated these nesting traits of dinosaurs, and the findings of each chapter are summarized below. Chapter 2 aimed to evaluate the accuracy of the previous method that is used to infer nest type of archosaur species, including dinosaurs. The previous method uses estimates of water vapor conductance for the eggs (i.e., diffusive capacity of eggshell: GH2O), which is experimentally measured for extant species and is calculated based on eggshell morphometrics for fossil eggs. Because experimental GH2O values for extant archosaur species appear to reflect nest type, morphometric GH2O values of extinct archosaur species have been used to infer their nest type (i.e., covered vs. open). Thus, previous studies assumed that GH2O values derived from the two methods were comparable, although this assumption has not been statistically evaluated.

Comparisons between experimental and morphometric GH2O values revealed that there is a disagreement between these methods. The disagreement was particularly apparent in small eggs, likely due to systematic errors. These results suggested that morphometric and experimental

GH2O of extant species are not necessarily comparable, although the reason for the discrepancy remains uncertain. Thus, direct comparisons between morphometric GH2O of dinosaurs and experimental GH2O of extant species should be avoided when inferring the nest type for dinosaurs. Considering the results of Chapter 2, Chapter 3 proposed an alternative approach to infer nest type of dinosaurs based on the eggshell porosity and egg mass. The dataset of extant archosaur species revealed a strong correlation between eggshell porosity and nest type, indicating that eggshell porosity can be used as a proxy for nest type (i.e., covered vs. open).

130

Prediction of nest type for extinct archosaurs suggested that: 1) covered nests are likely the primitive condition for dinosaurs (and probably archosaurs), and 2) open nests first evolved among non-avian theropods more derived than Lourinhanosaurus and were likely widespread in non-avian maniraptorans (well before the appearance of birds). Combined with taphonomic evidence, it can be suggested that basal open nesters (i.e., oviraptorosaurs and troodontids) were potentially the first dinosaurs to brood their clutches, but they still partially buried their eggs in sediment. A potential co-evolution of open nests and brooding behavior among maniraptorans may have freed theropods from the ground-based restrictions inherent to covered nests and allowed the exploitation of alternate nesting locations. These changes in nesting styles and behaviors thus may have played a role in the evolutionary success of maniraptorans (including birds). Chapter 4 aimed to infer nest structure and incubation heat sources of dinosaurs in order to discuss the paleogeography of dinosaur nesting sites, focusing on high latitudes, in particular. Unlike brooding birds, extant crocodylians and megapodes obtain incubation heat from surrounding environments for their covered nests. These species build either mound nests that are usually composed of plant materials and/or soil and obtain incubation heat from microbial respiration, or in-filled hole nests that are often composed of sand and obtain heat from solar radiation or geothermal activities. Incubation heat derived from microbial respiration keeps adequate nest temperature even though ambient air temperature is as low as 16 °C. Lithologic data of dinosaur clutches suggest that sauropods (megaloolithid eggs) and hadrosaurs could have incubated their eggs in mound nests mainly using heat from microbial respiration, whereas another type of sauropods (faveoloolithid eggs) likely incubated the eggs in in-filled hole nests using heat from solar radiation or geothermal energy. Open nesters (e.g., oviraptorosaurs and troodontids) potentially provided heat to their eggs via brooding. Thus, it could be interpreted that dinosaur eggs known from northern latitudes were incubated using heat from microbial respiration (hadrosaurs) or from brooding (troodontids). Chapter 5 attempts to estimate incubation period (i.e., the time interval from oviposition to hatching: measured in days) for dinosaurs. Although incubation period widely varies among extant archosaur species and egg mass is not the sole factor affecting incubation period, multiple regressions of various incubation traits such as eggshell porosity, pore density, and nest type, as well as egg mass, are used to estimate incubation period for archosaurs with statistical

131

confidence. As a general trend, estimates of incubation period relative to egg mass for non-avian dinosaurs were shorter than that of crocodylians, but were comparable to that of extant bird species. Thus, it can be suggested that a major reduction of incubation period should have occurred in clades between Crurotarsi and Dinosauria. Relatively short incubation period along with a large clutch size in dinosaurs could have been selected as an offset to high nest predation pressure. Findings in this dissertation along with previous studies allow discussion of reproductive style in archosaurs. Although some of nesting traits in crocodylians, ornithischians, sauropodomorphs, and Lourinhanosaurus are primitive among archosaurs (e.g., covered nest type, relatively small egg size, large clutch size, and precocial hatching), relatively short incubation could be inferred in Dinosauria. Features of more derived theropods, including oviraptorosaurs and troodontids, are characterized by monoautochronic ovulation (paired eggs) and asymmetrical eggs laid in a partially open nest, likely in association with brooding behavior. A shift from relatively small eggs with a large clutch size to relatively large eggs with a small clutch size could have occurred within maniraptorans, including Aves. Thus, it can be suggested that reproductive style and hence nesting strategy differs among major archosaurian clades and that a transition to more bird-like nesting methods occurred through archosaur evolution.

132

APPENDICES

Appendix 2.1. Museum specimens assessed for eggshell porosity in this study. Abbreviations of institutions: ROM, Royal Ontario Museum, Toronto, Canada; UHR, Hokkaido University Museum, Sapporo, Japan; YPM R., Herpetology Collection at the Yale Peabody Museum, New Haven, Connecticut; ZEC; Zelenitsky Egg Catalogue, the University of Calgary, Calgary, Canada. Taxon Collection number Aechmophorus occidentalis ROM 4693 Aix galericulata UHR 33131 Aix sponsa UHR 33128 Alectura lathami ZEC 137-1-3; ZEC 137-1-4; ZEC 137-1-5 Alligator mississippiensis YPM R. 015109; YPM R. 015110 Ammoperdix heyi ROM 9516 Anas bahamensis ROM 9487 Anas discors ROM 7563 Anas platyrhynchos ROM 12821 Anhinga anhinga ROM 9438 Anser anser ZEC 221-1-2; ZEC 221-1-3 Branta canadensis ZEC 444-1-1; ZEC 444-1-2 Bucephala islandica ROM 7739 Burhinus oedicnemus ROM 10983 Buteo rufinus ROM 8126 Cairina moschata ZEC 290-1-1; ZEC 290-1-2 Chlidonias niger ROM 10266 Chrysolophus amherstiae ROM 3659 Clangula hyemalis ROM 12427 Crocodylus porosus UHR 33210; YPM R. 17954; YPM R. 17976 Egretta thula ROM 4772 Egretta tricolor ROM 4782 Eudocimus albus ROM 4818

133

Taxon Collection number Eudromia elegans ZEC 283-1-2 Falco naumanni ROM 5096 Falco tinnunculus ROM 10855 Fratercula arctica ROM 2354 Fulmarus glacialis ROM 3034 Gavia immer ROM 13041 Larus glaucescens ROM 5503 Larus heermanni ROM 8686 Larus ridibundus ROM 3780 Leipoa ocellata ZEC 218-2-1; ZEC218-2-2 Lophura nycthemera ROM 4622 Nycticorax nycticorax UHR 33121 Oceanodroma leucorhoa ROM 2796 Onychoprion fuscatus ROM 5632 Oxyura jamaicensis ROM 8088 Passer domesticus ROM 2377; ZEC 455-1-1; ZEC 455-1-2 Pavo cristatus UHR 33126 Phalacrocorax pelagicus ZEC 445-1-1; ZEC 445-1-2 Plegadis falcinellus ROM 8002 cristatus ROM 1182 Podilymbus podiceps ROM 12592 Pygoscelis adeliae ROM 11421 Rissa tridactyla ROM 356 Rynchops niger ROM 10003 Somateria mollissima ROM 10863 Spheniscus demersus ROM 9973 Sterna hirundo ROM 8478 Sterna paradisaea ROM 5615 Sternula albifrons ROM 10864 Streptopelia turtur ROM 3665

134

Taxon Collection number Strix aluco ROM 3767 Syrmaticus soemmerringii UHR 33127

Tadorna tadorna ROM 9527 Thalasseus elegans ROM 13086 Thalasseus maximus ROM 5660 Tyto alba ROM 12631

135

Appendix 2.2. Mean values of egg mass (M), experimental water vapor conductance (Gex), –1 and eggshell porosity (Ap∙Ls ) for 106 living archosaur taxa.

Schonwetter (1960–1967) was used for egg length and breadth. *A value of Gex for Crocodylus porosus was calculated here based on daily water loss of eggs and known humidity from Grigg and Beard (1985). –1 Taxon M Gex Ap∙Ls References Accipitridae Buteo rufinus 60.720 6.260 1.027 Ar and Rahn (1985); this study Anseriformes Aix galericulata 39.750 5.483 3.952 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985); this study Aix sponsa 43.550 6.638 4.579 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1978, 1985); Hoyt et al. (1979); French and Board (1983); this study Anas bahamensis 34.450 8.500 1.989 Tullett (1976); Hoyt et al. (1979); Ar and Rahn (1985); this study Anas discors 25.400 4.600 0.550 Schonwetter (1960–1967); Hoyt et al. (1979); this study Anas fulvigula 55.950 16.300 11.246 Hoyt et al. (1979); Ar and Rahn (1985) Anas gracilis 33.850 7.700 3.368 Hoyt et al. (1979); French and Board (1983) Anas platyrhynchos 83.729 15.533 3.673 Schonwetter (1960–1967); Ar et al. (1974); Ar and Rahn (1978, 1985); Rokitka and Rahn (1987); Balkan et al. (2006); this study Anser anser 165.522 30.216 6.483 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1978, 1985);

136

–1 Taxon M Gex Ap∙Ls References Hoyt et al. (1979); Vleck et al. (1979); Rokitka and Rahn (1987); this study Anser 131.800 24.900 8.113 Hoyt et al. (1979); Rahn et al. brachyrhynchus (1983); Ar and Rahn (1985) Anser cygnoides 147.200 30.800 18.353 Hoyt et al. (1979); Ar and Rahn (1985) Anser erythropus 127.300 22.200 7.842 Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Anser fabalis 159.650 23.100 6.439 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); Ar and Rahn (1985) Branta canadensis 157.225 30.600 13.730 Schonwetter (1960–1967); Tullett (1976); Snyder et al. (1982); Ar and Rahn (1985); this study Branta h. minima 100.700 17.350 2.416 Hoyt et al. (1979); Ar and Rahn (1985) Branta leucopsis 108.900 21.700 4.216 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); French and Board (1983); Rahn et al. (1983); Ar and Rahn (1985) Branta sandvicensis 147.800 33.967 8.097 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Bucephala islandica 68.733 10.633 6.659 Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985); this study Cairina moschata 78.350 12.230 5.428 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn

137

–1 Taxon M Gex Ap∙Ls References (1978, 1985); French and Board (1983); this study Cereopsis 125.650 17.273 2.334 Schonwetter (1960–1967); Tullett novaehollandiae (1976); Ar and Rahn (1985); Wagner and Seymour (2001) Chloephaga 111.940 11.890 5.342 Carey et al. (1990) melanoptera Clangula hyemalis 45.900 11.600 1.885 Rahn et al. (1983); Ar and Rahn (1985); this study Cyanochen 83.600 15.033 4.604 Hoyt et al. (1979); French and cyanoptera Board (1983); Ar and Rahn (1985) Dendrocygna arborea 59.600 14.200 5.041 Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Dendrocygna 43.267 11.567 2.248 Hoyt et al. (1979); French and autumnalis Board (1983); Ar and Rahn (1985) Dendrocygna bicolor 53.467 16.433 3.868 Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Lophodytes cucullatus 55.750 8.700 1.806 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Mergus merganser 69.150 14.950 7.050 Hoyt et al. (1979); Ar and Rahn (1985) Mergus serrator 68.500 5.933 5.152 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Oxyura jamaicensis 73.467 20.233 13.167 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); French and Board (1983); Ar and Rahn

138

–1 Taxon M Gex Ap∙Ls References (1985); this study Oxyura leucocephala 95.200 22.067 10.940 Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Somateria mollissima 106.220 21.160 7.527 Schonwetter (1960–1967); Ar and Rahn (1978, 1985); Hoyt et al. (1979); French and Board (1983); Rahn et al. (1983); this study Tadorna tadorna 78.633 13.633 7.444 Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985); this study Tadorna variegate 90.867 13.500 12.231 Schonwetter (1960–1967); Tullett (1976); Hoyt et al. (1979); French and Board (1983); Ar and Rahn (1985) Charadriiformes Burhinus oedicnemus 33.500 4.575 4.800 Ar and Rahn (1978, 1985); this study Chlidonias niger 10.610 3.745 2.254 Schonwetter (1960–1967); Ar and Rahn (1985); Davis and Ackerman (1985); this study Fratercula arctica 59.783 7.993 4.777 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985); this study Fratercula cirrhata 89.933 13.000 4.956 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1978, 1985); Zimmermann and Hipfner (2007) Haematopus 41.483 6.800 1.750 Schonwetter (1960–1967); Ar et al. ostralegus (1974); Tullett (1976); Ar and Rahn (1978, 1985) Larus argentatus 88.487 16.813 4.727 Schonwetter (1960–1967); Ar et al.

139

–1 Taxon M Gex Ap∙Ls References (1974); Tullett (1976); Ar and Rahn (1978, 1985) Larus canus 58.438 13.481 3.946 Schonwetter (1960–1967); Ar et al. (1974); Lomholt (1976); Tullett (1976); Ar and Rahn (1978, 1985) Larus fuscus 84.600 15.943 9.464 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985) Larus glaucescens 97.400 23.187 7.574 Ar and Rahn (1978, 1985); Morgan et al. (1978); this study Larus heermanni 53.600 10.633 4.008 Ar and Rahn (1978, 1985); Rahn and Dawson (1979); this study Larus marinus 112.267 16.613 7.548 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985) Larus ridibundus 35.617 8.535 1.921 Lomholt (1976); Ar and Rahn (1978, 1985); this study Numenius phaeopus 53.487 9.727 5.946 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985) Onychoprion fuscatus 34.533 6.773 2.498 Rahn et al. (1976); Ar and Rahn (1978, 1985); this study Pluvialis apricaria 32.613 5.013 1.967 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985) Ptychoramphus 29.313 4.090 0.660 Schonwetter (1960–1967); aleuticus Roudybush et al. (1980); Ar and Rahn (1985); Zimmermann and Hipfner (2007) Rissa tridactyla 50.680 9.713 2.395 Schonwetter (1960–1967); Ar and

140

–1 Taxon M Gex Ap∙Ls References Rahn (1978, 1985); Morgan et al. (1978); this study Rynchops niger 26.450 6.800 1.105 Grant et al. (1984); Ar and Rahn (1985); this study Stercorarius skua 95.490 18.407 5.849 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985) Sterna hirundo 20.567 4.000 1.929 Rahn et al. (1976); Ar and Rahn (1978, 1985); this study Sterna paradisaea 18.300 5.073 0.488 Rahn et al. (1976); Ar and Rahn (1978, 1985); this study Sternula albifrons 8.790 1.907 0.415 Rahn et al. (1976); Ar and Rahn (1978, 1985); this study Thalasseus elegans 40.950 9.915 1.899 Ar and Rahn (1978, 1985); this study Thalasseus maximus 68.467 13.933 5.523 Schonwetter (1960–1967); Ar and Rahn (1978, 1985); Vleck et al. (1983); Rokitka and Rahn (1987); this study Uria aalge 114.600 20.050 4.995 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1978, 1985); Zimmermann and Hipfner (2007) Ciconiiformes Egretta thula 22.600 3.850 2.151 Vleck et al. (1983); Ar and Rahn (1985); this study Egretta tricolor 26.050 3.900 1.237 Vleck et al. (1983); Ar and Rahn (1985); this study Eudocimus albus 50.250 6.650 1.634 Vleck et al. (1983); Ar and Rahn (1985); this study Nycticorax nycticorax 33.700 6.300 1.458 Ar and Rahn (1978, 1985); Vleck et

141

–1 Taxon M Gex Ap∙Ls References al. (1983); this study Plegadis falcinellus 37.300 7.795 2.654 Vleck et al. (1983); Ar and Rahn (1985); this study Columbiformes Columba livia 18.428 4.448 1.121 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1978); Vleck et al. (1979); Arad et al. (1988) Streptopelia turtur 8.300 2.205 0.323 Ar and Rahn (1978, 1985); this study Falco naumanni 10.820 2.800 0.639 Ar and Rahn (1978, 1985); this study Falco tinnunculus 18.100 3.980 0.477 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1978); this study Galliformes Alectura lathami 192.500 47.167 32.757 Seymour and Rahn (1978); Ar and Rahn (1985); Seymour et al. (1987); Booth and Thompson (1991); this study Ammoperdix heyi 14.240 1.180 0.223 Ar and Rahn (1985); this study Chrysolophus 30.080 5.945 5.161 Ar and Rahn (1978, 1985); this amherstiae study Coturnix coturnix 10.038 3.054 1.416 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985); Vleck et al. (1979); Booth and Rahn (1989) Gallus gallus 56.611 13.823 1.949 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Rahn et al. (1977b); Ar and Rahn (1978, 1985);

142

–1 Taxon M Gex Ap∙Ls References Arad and Marder (1981); Tullett (1981); Tullett and Deeming (1982); Burton and Tullett (1983); Leon-Velarde et al. (1984); Visschedijk et al. (1985); Andersen and Steen (1986); Rokitka and Rahn (1987); Seymour and Visschedijk (1988); Booth and Rahn (1989); O'dea et al. (2004); Hamidu et al. (2007) Leipoa ocellata 173.000 43.500 46.814 Schonwetter (1960–1967); Tullett (1976); Vleck et al. (1984); Seymour et al. (1986, 1987); Booth and Seymour (1987); this study Lophophorus 64.555 8.350 2.091 Schonwetter (1960–1967); Tullett impejanus (1976); Ar and Rahn (1978, 1985) Lophura nycthemera 39.913 9.237 7.455 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985); this study Meleagris gallopavo 85.222 14.778 5.118 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976, 1981); Ar and Rahn (1978, 1985); Burton and Tullett (1983); Rokitka and Rahn (1987) Numida meleagris 48.900 10.000 1.658 Schonwetter (1960–1967); Tullett (1976); Ancel and Girard (1992) Pavo cristatus 95.085 14.020 13.626 Ar and Rahn (1978, 1985); this study Phasianus colchicus 32.280 6.813 2.203 Schonwetter (1960–1967); Ar et al. (1974); Tullett (1976); Ar and Rahn

143

–1 Taxon M Gex Ap∙Ls References (1978, 1985) Syrmaticus 31.475 7.925 2.622 Ar and Rahn (1978, 1985); this soemmerringii study Gavia immer 154.000 98.000 18.679 Schonwetter (1960–1967); Tullett (1976); Tullett and Board (1977); Ar and Rahn (1985); This study. Passeriformes Agelaius phoeniceus 4.062 1.415 0.042 Ar et al. (1974); Rahn et al. (1977b); Ar and Rahn (1978); Carey (1979); Carey et al. (1983); Jaeckle et al. (2012) Menura 60.300 7.080 1.107 Lill (1987); Higgins et al. (2001) novaehollandiae Molothrus ater 3.330 1.070 0.065 Ar et al. (1974); Ar and Rahn (1978); Baicich and Harrison (1997); Jaeckle et al. (2012) Passer domesticus 2.698 0.825 0.136 Ar et al. (1974); Tullett (1976); Ar and Rahn (1978, 1985); this study Turdus merula 6.607 1.750 0.514 Tullett (1976); Ar and Rahn (1978, 1985); Higgins et al. (2006) Anhinga anhinga 36.340 6.120 2.937 Ar and Rahn (1985); Colacino et al. (1985); this study Phalacrocorax 52.427 6.420 2.841 Ar et al. (1974); Tullett (1976); Ar auritus and Rahn (1978, 1985) Phalacrocorax 39.450 6.770 2.868 Ar and Rahn (1978, 1985); this pelagicus study Podicipediformes Aechmophorus 49.370 21.640 17.547 Ar and Rahn (1985); this study

144

–1 Taxon M Gex Ap∙Ls References occidentalis Podiceps cristatus 41.100 19.480 11.162 Schonwetter (1960–1967); Lomholt (1976); this study Podilymbus podiceps 21.063 12.770 9.338 Schonwetter (1960–1967); Ackerman and Platter-Rieger (1979); Davis et al. (1984); Ar and Rahn (1985); this study Procellariiformes Fulmarus glacialis 100.900 12.400 13.346 Rahn et al. (1984); Ar and Rahn (1985); this study Oceanodroma 10.500 1.550 0.507 Ar and Rahn (1985); Rahn and leucorhoa Huntington (1988); this study pacificus 58.793 6.220 6.698 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1978, 1985); Ackerman et al. (1980) Puffinus tenuirostris 87.400 10.650 8.462 Schonwetter (1960–1967); Fitzherbert (1985) Sphenisciformes Aptenodytes 302.300 26.650 14.192 Schonwetter (1960–1967); Tullett patagonicus (1976); Ar and Rahn (1985); Handrich (1989) Pygoscelis adeliae 122.300 14.033 8.126 Rahn and Hammel (1982); Ar and Rahn (1985); Thompson and Goldie (1990); this study Spheniscus demersus 100.750 15.000 4.920 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1985); this study Strigiformes Strix aluco 36.100 6.175 1.551 Ar and Rahn (1978, 1985); this study

145

–1 Taxon M Gex Ap∙Ls References Tyto alba 18.700 5.100 1.336 Ar and Rahn (1985); this study Struthioniformes Apteryx australis 350.200 26.000 10.120 Schonwetter (1960–1967); Calder (1978); Calder et al. (1978); Silyn-Roberts (1983) Tinamiformes Eudromia elegans 35.800 7.750 2.561 Ar and Rahn (1985); this study Crocodylia Alligator 72.270 51.593 31.385 Packard et al. (1979); Wink et al. mississippiensis (1990); Marzola et al. (2015); this study Crocodylus porosus 95.652 31.836* 21.446 Ferguson (1985); Grigg and Beard (1985); this study

146

Appendix 2.3. Log-scaled variables of water vapor conductance (GH2O) of eggs for statistical analyses.

Log Gex and log Gmo represent experimental and morphometric log-transformed GH2O values, respectively. Difference and mean were calculated from log Gex and log Gmo (see "2.2 Materials and Methods").

Taxon log M log Gex log Gmo Difference Mean Accipitridae Buteo rufinus 1.783 0.797 0.334 0.463 0.565 Anseriformes Aix galericulata 1.599 0.739 0.919 -0.180 0.829 Aix sponsa 1.639 0.822 0.983 -0.161 0.903 Anas bahamensis 1.537 0.929 0.621 0.309 0.775 Anas discors 1.405 0.663 0.063 0.600 0.363 Anas fulvigula 1.748 1.212 1.373 -0.161 1.293 Anas gracilis 1.530 0.886 0.850 0.037 0.868 Anas platyrhynchos 1.923 1.191 0.887 0.304 1.039 Anser anser 2.219 1.480 1.134 0.346 1.307 Anser brachyrhynchus 2.120 1.396 1.231 0.165 1.314 Anser cygnoides 2.168 1.489 1.586 -0.097 1.537 Anser erythropus 2.105 1.346 1.217 0.130 1.281 Anser fabalis 2.203 1.364 1.131 0.233 1.247 Branta canadensis 2.197 1.486 1.460 0.026 1.473 Branta h. minima 2.003 1.239 0.705 0.534 0.972 Branta leucopsis 2.037 1.336 0.947 0.389 1.142 Branta sandvicensis 2.170 1.531 1.231 0.300 1.381 Bucephala islandica 1.837 1.027 1.146 -0.119 1.086 Cairina moschata 1.894 1.087 1.057 0.031 1.072 Cereopsis novaehollandiae 2.099 1.237 0.690 0.547 0.964 Chloephaga melanoptera 2.049 1.075 1.050 0.025 1.063 Clangula hyemalis 1.662 1.064 0.597 0.467 0.831 Cyanochen cyanoptera 1.922 1.177 0.985 0.192 1.081

147

Taxon log M log Gex log Gmo Difference Mean Dendrocygna arborea 1.775 1.152 1.025 0.128 1.089 Dendrocygna autumnalis 1.636 1.063 0.674 0.389 0.869 Dendrocygna bicolor 1.728 1.216 0.910 0.306 1.063 Lophodytes cucullatus 1.746 0.940 0.579 0.361 0.759 Mergus merganser 1.840 1.175 1.170 0.004 1.173 Mergus serrator 1.836 0.773 1.034 -0.261 0.904 Oxyura jamaicensis 1.866 1.306 1.442 -0.136 1.374 Oxyura leucocephala 1.979 1.344 1.361 -0.017 1.352 Somateria mollissima 2.026 1.326 1.199 0.127 1.262 Tadorna tadorna 1.896 1.135 1.194 -0.059 1.164 Tadorna variegata 1.958 1.130 1.410 -0.279 1.270 Charadriiformes Burhinus oedicnemus 1.525 0.660 1.003 -0.343 0.832 Chlidonias niger 1.026 0.573 0.675 -0.102 0.624 Fratercula arctica 1.777 0.903 1.001 -0.099 0.952 Fratercula cirrhata 1.954 1.114 1.017 0.097 1.066 Haematopus ostralegus 1.618 0.833 0.565 0.267 0.699 Larus argentatus 1.947 1.226 0.997 0.229 1.111 Larus canus 1.767 1.130 0.918 0.211 1.024 Larus fuscus 1.927 1.203 1.298 -0.096 1.250 Larus glaucescens 1.989 1.365 1.202 0.164 1.283 Larus heermanni 1.729 1.027 0.925 0.102 0.976 Larus marinus 2.050 1.220 1.200 0.020 1.210 Larus ridibundus 1.552 0.931 0.606 0.325 0.769 Numenius phaeopus 1.728 0.988 1.096 -0.109 1.042 Onychoprion fuscatus 1.538 0.831 0.720 0.111 0.775 Pluvialis apricaria 1.513 0.700 0.616 0.084 0.658 Ptychoramphus aleuticus 1.467 0.612 0.142 0.470 0.377 Rissa tridactyla 1.705 0.987 0.701 0.286 0.844 Rynchops niger 1.422 0.833 0.365 0.467 0.599

148

Taxon log M log Gex log Gmo Difference Mean Stercorarius skua 1.980 1.265 1.089 0.176 1.177 Sterna hirundo 1.313 0.602 0.608 -0.005 0.605 Sterna paradisaea 1.262 0.705 0.011 0.695 0.358 Sternula albifrons 0.944 0.280 -0.060 0.340 0.110 Thalasseus elegans 1.612 0.996 0.601 0.395 0.799 Thalasseus maximus 1.835 1.144 1.064 0.080 1.104 Uria aalge 2.059 1.302 1.021 0.281 1.161 Ciconiiformes Egretta thula 1.354 0.585 0.655 -0.069 0.620 Egretta tricolor 1.416 0.591 0.414 0.177 0.503 Eudocimus albus 1.701 0.823 0.535 0.287 0.679 Nycticorax nycticorax 1.528 0.799 0.486 0.313 0.643 Plegadis falcinellus 1.572 0.892 0.746 0.146 0.819 Columbiformes Columba livia 1.265 0.648 0.372 0.276 0.510 Streptopelia turtur 0.919 0.343 -0.169 0.512 0.087 Falconidae Falco naumanni 1.034 0.447 0.128 0.319 0.288 Falco tinnunculus 1.258 0.600 0.001 0.599 0.300 Galliformes Alectura lathami 2.284 1.674 1.838 -0.164 1.756 Ammoperdix heyi 1.154 0.072 -0.330 0.402 -0.129 Chrysolophus amherstiae 1.478 0.774 1.035 -0.261 0.905 Coturnix coturnix 1.002 0.485 0.473 0.012 0.479 Gallus gallus 1.753 1.141 0.612 0.529 0.876 Leipoa ocellata 2.238 1.638 1.993 -0.354 1.816 Lophophorus impejanus 1.810 0.922 0.643 0.279 0.782 Lophura nycthemera 1.601 0.966 1.195 -0.229 1.080 Meleagris gallopavo 1.931 1.170 1.031 0.138 1.100 Numida meleagris 1.689 1.000 0.542 0.458 0.771

149

Taxon log M log Gex log Gmo Difference Mean Pavo cristatus 1.978 1.147 1.457 -0.310 1.302 Phasianus colchicus 1.509 0.833 0.665 0.168 0.749 Syrmaticus soemmerringii 1.498 0.899 0.741 0.158 0.820 Gaviiformes Gavia immer 2.188 1.991 1.594 0.398 1.792 Passeriformes Agelaius phoeniceus 0.609 0.151 -1.051 1.201 -0.450 Menura novaehollandiae 1.780 0.850 0.366 0.484 0.608 Molothrus ater 0.522 0.029 -0.868 0.897 -0.419 Passer domesticus 0.431 -0.084 -0.544 0.460 -0.314 Turdus merula 0.820 0.243 0.033 0.210 0.138 Pelecaniformes Anhinga anhinga 1.560 0.787 0.790 -0.003 0.788 Phalacrocorax auritus 1.720 0.808 0.776 0.032 0.792 Phalacrocorax pelagicus 1.596 0.831 0.780 0.051 0.805 Podicipediformes Aechmophorus occidentalis 1.693 1.335 1.566 -0.231 1.451 Podiceps cristatus 1.614 1.290 1.370 -0.080 1.330 Podilymbus podiceps 1.324 1.106 1.292 -0.186 1.199 Procellariiformes Fulmarus glacialis 2.004 1.093 1.448 -0.354 1.270 Oceanodroma leucorhoa 1.021 0.190 0.027 0.163 0.109 Puffinus pacificus 1.769 0.794 1.148 -0.354 0.971 Puffinus tenuirostris 1.942 1.027 1.250 -0.222 1.139 Sphenisciformes Aptenodytes patagonicus 2.480 1.426 1.474 -0.049 1.450 Pygoscelis adeliae 2.087 1.147 1.232 -0.085 1.190 Spheniscus demersus 2.003 1.176 1.014 0.162 1.095 Strigiformes Strix aluco 1.558 0.791 0.513 0.278 0.652

150

Taxon log M log Gex log Gmo Difference Mean Tyto alba 1.272 0.708 0.448 0.260 0.578 Struthioniformes Apteryx australis 2.544 1.415 1.327 0.088 1.371 Tinamiformes Eudromia elegans 1.554 0.889 0.731 0.159 0.810 Crocodylia Alligator mississippiensis 1.859 1.713 1.819 -0.106 1.766 Crocodylus porosus 1.981 1.503 1.654 -0.151 1.578

151

Appendix 2.4. Cladogram of 106 living archosaur taxa that were used for independent contrast method.

Tree sources from Kennedy and Spencer (2000), Kennedy and Page (2002), Fjeldsa (2004),

Thomas et al. (2004), Crowe et al. (2006), Ksepka et al. (2006), Hackett et al. (2008), McCracken et al. (2010), Gibb et al. (2013), and Ramirez et al. (2013).

152

Appendix 3.1. Use of eggshell porosity as an indicator of nest types in archosaurs.

−1 It can be hypothesized that values of eggshell porosity (Ap∙Ls ) differ between species with covered and open nests in living archosaurs. Because pore canals are the only pathways for water vapor in eggshell, water vapor conductance (GH2O) can be calculated using eggshell properties

(morphometric method), which is deduced from Fick's first law of gas diffusion:

[A3.1-1]

−1 −1 where c is a unit conversion constant (mg H2O∙sec∙day ∙mol ), DH2O is the diffusion coefficient

2 −1 −1 −1 of water vapor (mm ∙sec ), Gmo is the morphometrically-derived GH2O (mg H2O∙day ∙Torr ), R is the universal gas constant (mm3∙Torr∙mol−1∙K−1), and T is the absolute temperature of incubation

(°K) (Ar et al., 1974). Because DH2O, T, c, and R are assumed to be consistent among species (e.g.,

Ar et al., 1974; Ar and Rahn, 1985; Tanaka and Zelenitsky, 2014a), Equation A3.1-1 can be simplified (Tanaka and Zelenitsky, 2014a):

[A3.1-2]

Also, GH2O is experimentally estimated (experimental method) using fresh eggs as:

[A3.1-3]

−1 where MH2O is the daily loss of water vapor (mg H2O∙day ), Gex is the experimentally-derived

−1 −1 GH2O (mg H2O∙day ∙Torr ), Pegg is the water vapor pressure inside of the egg (Torr), and Pnest is the water vapor pressure of the nest (Torr) (Ar et al., 1974). Although log Gmo may not be equivalent to log Gex due to potential systematic errors between the morphometric and experimental methods, these variables show a significant positive correlation (Tanaka and

Zelenitsky, 2014a). Thus,

[A3.1-4] and 153

[A3.1-5]

−1 Thus, log Ap∙Ls is proportional to log MH2O and inversely proportional to log (Pegg – Pnest). While

Pegg is assumed to be consistent among species (40–50 Torr in birds: Booth and Thompson, 1991;

Rahn, 1984; and approximately 32 Torr in crocodylians: Tanaka and Zelenitsky, 2014b), Pnest is significantly higher in species with covered nests (mean 31.74 Torr) than in those with open nests

(mean 20.68 Torr) (Tanaka and Zelenitsky, 2014b). Thus, log (Pegg – Pnest) should be lower in species with covered nests, especially crocodilians, than in those with open nests. Also, log MH2O values, relative to log egg mass (M, in g), are consistent among species, regardless of nest types

(Appendices 3.2 and 3.7), because no outlier was detected based on a Generalized Extreme

Studentized Deviate test using MedCalc Statistical Software v. 13.0 (MedCalc Software bvba,

−1 Ostend, Belgium; http://www.medcalc.org). Therefore, it can be assumed that Ap∙Ls is primarily

−1 affected by Pnest and that log Ap∙Ls is expected to be higher in species that build covered nests

(i.e., high Pnest) than in those that build open nests (i.e., lower Pnest).

154

Appendix 3.2. Daily loss of water vapor (MH2O) and egg mass (M) in 196 living archosaur species.

Values of M and MH2O, both derived from natural incubation experiments, were compiled from the literature. Some MH2O values were calculated in the context of this study based on values reported in the literature of mass loss of eggs over known time intervals. When multiple values of log MH2O and log M were available for a species, mean values were calculated.

Taxon log MH2O log M References

Accipitridae

Accipiter cooperii 2.328 1.462 Sumner (1929b)

Accipiter fasciatus 2.114 1.583 Marchant and Higgins (1994)

Aquila chrysaetos 2.642 2.161 Sumner (1929a, 1929b)

Buteo buteo 2.398 1.797 Groebbles (1927); Drent

(1970)

Rostrhamus sociabilis 2.292 1.501 Sykes (1987)

Anseriformes

Anas puna 2.595 1.659 Carey et al. (1989b)

Anser a. frontalis 2.845 2.107 Ely and Raveling (1984); Ely

and Dzubin (1994)

Anser anser 2.869 2.241 Ar and Rahn (1985)

Bucephala clangula 2.428 1.806 Eadie et al. (1995)

Bucephala islandica 2.398 1.829 Eadie et al. (2000)

Cygnus olor 3.170 2.542 Booth and Sotherland (1991)

155

Taxon log MH2O log M References

Lophodytes cucullatus 2.517 1.768 Dugger et al. (1994)

Somateria mollissima 2.812 2.020 Gross (1938); Belopol'skii

(1961); Drent (1970);

Hagelund and Norderhaug

(1973)

Apodiformes

Cypseloides niger 1.672 0.722 Lowther and Collins (2002)

Caprimulgiformes

Chordeiles acutipennis 1.598 0.760 Grant (1982)

Charadriiformes

Actitis macularius 1.820 0.982 Lank et al. (1985)

Alca torda 2.556 1.964 Plumb (1965); Drent (1970)

Alle alle 2.314 1.491 Stempniewicz (1981); Rahn

and Paganelli (1990)

Anous minutus marcusi 2.025 1.375 Rahn et al. (1976)

Anous stolidus 2.102 1.548 Rahn et al. (1976); Morris and

Chardine (1992)

Anous tenuirostris 2.025 1.394 Whittow (1980)

Brachyramphus brevirostris 2.477 1.732 Andreev and Golubova (1995)

Calidris himantopus 2.461 1.683 Klima and Jehl (1998)

Cepphus grylle 2.408 1.706 Belopol'skii (1961); Drent

(1970); Asbirk (1979)

156

Taxon log MH2O log M References

Charadrius alexandrinus 1.579 0.874 Grant (1982)

Charadrius vociferus 1.835 1.159 Ar and Rahn (1980); Manning

(1981); Grant (1982);

Chlidonias niger 1.851 1.026 Davis and Ackerman (1985)

Fratercula arctica 2.469 1.831 Belopol'skii (1961); Drent

(1970)

Fratercula cirrhata 2.482 1.959 Piatt and Kitaysky (2002b)

Fratercula corniculata 2.534 1.878 Piatt and Kitaysky (2002a)

Gallinago gallinago 2.114 1.220 Manning (1981)

Gelochelidon nilotica 2.068 1.455 Grant et al. (1984)

Gygis alba 1.883 1.349 Rahn et al. (1976); Pettit et al.

(1981)

Haematopus bachmani 2.398 1.663 Andres and Falxa (1995)

Haematopus ostralegus 2.505 1.638 Barth (1953)

Himantopus mexicanus 2.180 1.322 Grant (1982)

Larus argentatus 2.721 1.962 Barth (1953); Belopol'skii

(1961); Drent (1970); Harris

(1964); Ar and Rahn (1985)

Larus atricilla 2.362 1.649 Vleck et al. (1983)

Larus canus 2.556 1.715 Barth (1953); Belopol'skii

(1961); Drent (1970)

Larus dominicanus 3.087 1.929 Fordham (1964a, 1964b)

157

Taxon log MH2O log M References

Larus fuscus 2.568 1.885 Barth (1953)

Larus glaucescens 2.753 1.992 Rahn et al. (1977a); Morgan et

al. (1978)

Larus heermanni 2.444 1.728 Rahn and Dawson (1979)

Larus livens 2.744 1.990 Rahn and Dawson (1979)

Larus marinus 2.744 2.042 Belopol'skii (1961); Harris

(1964); Drent (1970); Ar and

Rahn (1985)

Larus novaehollandiae 2.230 1.623 Wooller and Dunlop (1980)

Larus serranus 2.679 1.744 Carey et al. (1987)

Numenius phaeopus 2.575 1.688 Skeel and Mallory (1996)

Onychoprion fuscatus 2.279 1.553 Rahn et al. (1976)

Onychoprion lunatus 2.039 1.458 Whittow (1983); Whittow et

al. (1985)

Phalaropus fulicarius 1.818 0.898 Tracy et al. (2002)

Phalaropus tricolor 1.701 0.978 Manning (1981)

Pluvialis squatarola 2.176 1.491 Hussell and Page (1976)

Pluvianus aegyptius 1.544 0.968 Howell (1979)

Ptychoramphus aleuticus 2.004 1.473 Roudybush et al. (1980)

Recurvirostra americana 2.213 1.511 Grant (1982)

Rissa tridactyla 2.521 1.707 Maunder and Threlfall (1972);

Rahn et al. (1977a); Morgan et

158

Taxon log MH2O log M References

al. (1978); Rahn and Paganelli

(1990)

Rynchops niger 2.155 1.430 Grant et al. (1984)

Stercorarius longicaudus 2.431 1.604 Wiley and Lee (1998)

Stercorarius maccormicki 2.751 1.996 Rahn et al. (1977a); Ar and

Rahn (1980)

Stercorarius parasiticus 2.653 1.701 Belopol'skii (1961); Drent

(1970)

Sterna forsteri 1.903 1.328 Grant (1982); McNicholl et al.

2001)

Sterna hirundo 2.134 1.312 Barth (1953); Rahn et al.

(1976); Ar and Rahn (1985)

Sterna paradisaea 2.122 1.270 Belopol'skii (1961); Drent

(1970); Rahn et al. (1976)

Sternula albifrons 1.792 0.940 Rahn et al. (1976)

Thalasseus maximus 2.507 1.833 Vleck et al. (1983)

Thalasseus sandvicensis 2.279 1.558 Vleck et al. (1983)

Tringa semipalmatus 2.387 1.604 Ar and Rahn (1980); Raynor

and Wilcox (1980)

Tringa totanus 2.072 1.322 Barth (1953)

Uria lomvia 2.701 2.020 Uspenski (1958); Belopol'skii

(1961); Drent (1970); Gaston

159

Taxon log MH2O log M References

and Nettleship (1981); Ar and

Rahn (1985)

Vanellus miles 2.144 1.491 Marchant and Higgins (1994)

Ciconiiformes

Ardea albus 2.364 1.687 Vleck et al. (1983)

Ardea purpurea 2.591 1.699 Ar and Rahn (1980)

Botaurus lentiginosus 2.306 1.504 Manning (1981)

Bubulcus ibis 2.221 1.450 Ar and Rahn (1980); Vleck et

al. (1983)

Egretta garzetta 2.380 1.425 Ar and Rahn (1980)

Egretta thula 2.083 1.354 Vleck et al. (1983); Ar and

Rahn (1985)

Egretta tricolor 2.100 1.425 Vleck et al. (1983)

Eudocimus albus 2.462 1.706 Vleck et al. (1983)

Nycticorax nycticorax 2.261 1.519 Ar and Rahn (1980); Vleck et

al. (1983)

Plegadis falcinellus 2.316 1.573 Vleck et al. (1983)

Plegadis ridgwayi 2.481 1.530 Carey et al. (1987)

Coraciiformes

Merops ornatus 1.415 0.629 Lill and Fell (2007)

Falconidae

Falco cenchroides 1.699 1.276 Marchant and Higgins (1994)

160

Taxon log MH2O log M References

Falco sparverius 1.908 1.152 Sherman (1913); Drent (1970)

Falco tinnunculus 2.230 1.326 Groebbles and Mobert (1927);

Drent (1970)

Galliformes

Alectura lathami 2.531 2.255 Vleck et al. (1984); Seymour et

al. (1987)

Centrocercus urophasianus 2.322 1.580 Patterson (1952); Drent (1970)

Gallus gallus 2.687 1.752 Burke (1925); Chattock

(1925); Horton (1932)

Lagopus lagopus 2.193 1.328 Andersen and Steen (1986);

Steen et al. (1988)

Lagopus muta 2.262 1.372 Steen and Unander (1985)

Leipoa ocellata 2.425 2.238 Vleck et al. (1984); Seymour et

al. (1987)

Phasianus colchicus 2.308 1.491 Gladstone (1904); Groebbles

and Mobert (1927); Drent

(1970); Rahn et al. (1977a); Ar

and Rahn (1985)

Gaviiformes

Gavia stellata 2.791 1.887 Barr et al. (2000)

Gruiformes

Coturnicops noveboracensis 1.859 0.839 Elliot and Morrison (1979)

161

Taxon log MH2O log M References

Fulica americana 2.273 1.507 Gullion (1954); Davis et al.

(1984); Sotherland et al.

(1984); Carey et al. (1989a)

Gallinula tenebrosa 2.274 1.524 Lill (1990)

Grus canadensis 2.914 2.229 Walkinshaw (1950a); Drent

(1970)

Porphyrio porphyrio 2.267 1.546 Lill (1990)

Porzana carolina 1.724 0.939 Walkinshaw (1940); Manning

(1981)

Rallus elegans 2.041 1.279 Meanley (1969)

Passeriformes

Agelaius phoeniceus 1.704 0.609 Ar and Rahn (1980); Manning

(1981, 1982)

Aphelocoma coerulescens 1.662 0.764 Woolfenden (1978);

Woolfenden and Fitzpatrick

(1996)

Calcarius lapponicus 1.623 0.435 Hussell (1972); Hussell and

Montgomerie (2002)

Calcarius mccownii 1.699 0.380 Mickey (1943)

Carduelis cannabina 1.342 0.199 Groebbles and Mobert (1927);

Drent (1970)

Carduelis tristis 1.176 0.176 Ar and Rahn (1980)

162

Taxon log MH2O log M References

Corvus caurinus 2.287 1.250 Butler et al. (1984)

Delichon urbicum 1.301 0.281 Barth (1953)

Dendroica petechia 1.204 0.214 Schrantz (1943); Drent (1970);

Ar and Rahn (1980)

Dumetella carolinensis 1.690 0.568 Ar and Rahn (1980)

Emberiza citrinella 1.643 0.452 Groebbles and Mobert (1927);

Drent (1970)

Erithacus rubecula 1.531 0.382 Groebbles and Mobert (1927);

Drent (1970)

Euplectes orix 1.220 0.238 Woodall and Parry (1982)

Ficedula hypoleuca 1.257 0.220 Kern et al. (1992)

Fringilla coelebs 1.644 0.324 Groebbles and Mobert (1927);

Drent (1970)

Gerygone igata 1.313 0.173 Higgins and Peter (2002)

Heteromyias albispecularis 1.886 0.699 Frith and Frith (2000)

Hirundo rustica 1.293 0.302 Manning (1981, 1982)

Junco hyemalis 1.301 0.387 Nolan et al. (2002)

Lanius ludovicianus 1.721 0.667 Miller (1931)

Leucosticte australis 1.952 0.414 Johnson et al. (2000)

Melospiza melodia micronyx 1.442 0.455 Kern et al. (1990)

Menura novaehollandiae 2.288 1.789 Lill (1987)

Orthonyx spaldingii 1.851 1.152 Higgins and Peter (2002)

163

Taxon log MH2O log M References

Passer domesticus 1.506 0.463 Weaver (1943); Dawson

(1964); Ar and Rahn (1980);

Manning (1982)

Passer moabiticus 1.204 0.255 Ar and Rahn (1980)

Passerculus sandwichensis 1.524 0.338 Manning (1981)

Passerina cyanea 1.519 0.320 Morgan (1976)

Phylloscopus trochilus 1.097 -0.022 Barth (1953)

Plectrophenax nivalis 1.681 0.501 Hussell (1972)

Ploceus capensis 1.653 0.560 Brown (1994)

Prionodura newtoniana 2.090 1.061 Frith and Frith (1998)

Prunella modularis 1.477 0.328 Groebbles and Mobert (1927);

Drent (1970)

Ptiloris victoriae 2.015 1.009 Frith and Frith (1995)

Quiscalus major 1.921 0.908 Post et al. (1996)

Quiscalus quiscula 1.787 0.827 Manning (1981)

Sayornis nigricans 1.383 0.322 Wolf (1997)

Sayornis phoebe 1.301 0.312 Manning (1981, 1982)

Sialia sialis 1.486 0.464 Hamilton (1943)

Sturnus vulgaris 1.909 0.846 Manning (1982)

Sylvia communis 1.431 0.255 Groebbles and Mobert (1927);

Drent (1970)

Tachycineta bicolor 1.299 0.279 Manning (1982)

164

Taxon log MH2O log M References

Toxostoma rufum 1.819 0.740 Cavitt and Haas (2000)

Troglodytes aedon 1.216 0.165 Kendeigh (1940, 1963)

Turdus iliacus 1.826 0.777 Foster (1902); Drent (1970)

Turdus merula 1.886 0.839 Foster (1902); Drent (1970)

Turdus migratorius 1.843 0.805 Manning (1981, 1982)

Turdus philomelos 1.708 0.763 Groebbles and Mobert (1927);

Drent (1970)

Turdus pilaris 2.000 0.903 Barth (1953)

Xanthocephalus 1.739 0.653 Twedt and Crawford (1995) xanthocephalus

Xenicus gilviventris 1.557 0.407 Higgins et al. (2001)

Pelecaniformes

Fregata minor 2.309 1.950 Whittow (1983); Whittow et

al. (2003)

Morus bassanus 2.431 2.019 Nelson (1966)

Phalacrocorax aristotelis 2.366 1.683 Snow (1960); Belopol'skii

(1961); Drent (1970)

Sula abbotti 2.318 2.027 Nelson (1971, 2006)

Sula sula 2.249 1.766 Whittow (1980, 1983);

Whittow et al. (1989)

Phaethontidae

Phaethon aethereus 2.301 1.797 Nelson (2006)

165

Taxon log MH2O log M References

Phaethon lepturus 2.230 1.610 Stonehouse (1962)

Phaethon rubricauda 2.299 1.855 Whittow (1983)

Piciformes

Dendrocopos syriacus 1.908 0.732 Ar and Rahn (1985)

Podicipediformes

Podiceps nigricollis 2.245 1.340 Sotherland et al. (1984)

Podilymbus podiceps 2.126 1.327 Ackerman and Platter-Rieger

(1979); Davis et al. (1984)

Procellariiformes

Bulweria bulwerii 1.878 1.320 Whittow (1983); Robertson

and James (1988); Whittow

and Pettit (2000)

Calonectris diomedea borealis 2.483 1.986 Robertson and James (1988)

Diomedea exulans 3.024 2.707 Tickell (1968); Ar and Rahn

(1980); Brown and Adams

(1988)

Hydrobates pelagicus 1.602 0.845 Ar and Rahn (1980)

Macronectes giganteus 2.843 2.433 Brown and Adams (1988)

Oceanodroma castro 1.591 0.929 Robertson and James (1988)

Oceanodroma furcata 1.563 1.097 Boersma and Wheelwright

(1979)

Oceanodroma leucorhoa 1.666 1.023 Ar and Rahn (1980); Rahn and

166

Taxon log MH2O log M References

Huntington (1988)

Pachyptila vittata 2.068 1.544 Brown and Adams (1988)

Phoebastria immutabilis 2.852 2.449 Fisher (1969); Grant et al.

(1982c)

Phoebastria nigripes 2.849 2.484 Grant et al. (1982c)

Procellaria aequinoctialis 2.535 2.121 Brown and Adams (1988)

Procellaria westlandica 2.580 2.176 Baker and Coleman (1977)

Pterodroma alba 2.090 1.748 Rahn and Whittow (1988)

Pterodroma hypoleuca 2.000 1.593 Whittow (1980); Grant et al.

(1982b)

Pterodroma inexpectata 2.342 1.790 Whittow (1980)

Pterodroma phaeopygia 2.439 1.886 Whittow et al. (1984)

Puffinus nativitatis 2.110 1.695 Whittow (2001)

Puffinus pacificus 2.178 1.764 Ackerman et al. (1980); Ar and

Rahn (1980); Whittow et al.

(1982)

Puffinus puffinus 2.148 1.766 Harris (1966)

Puffinus tenuirostris 2.431 1.942 Fitzherbert (1985)

Sphenisciformes

Aptenodytes patagonicus 2.867 2.481 Ar and Rahn (1985); Handrich

(1989)

Spheniscus demersus 2.614 2.003 Ar and Rahn (1985)

167

Taxon log MH2O log M References

Pygoscelis adeliae 2.630 2.096 Rahn et al. (1977a); Rahn and

Hammel (1982)

Strigiformes

Aegolius funereus 1.908 1.101 Kuhk (1949); Drent (1970)

Asio otus 2.209 1.405 Sumner (1929b)

Bubo virginianus 2.284 1.729 Sumner (1929b)

Megascops asio 2.087 1.236 Sherman (1911); Sumner

(1928)

Tyto alba 2.079 1.326 Sumner (1929b); Drent (1970);

Ar and Rahn (1985)

Struthioniformes

Apteryx australis 2.815 2.615 Colbourne (2002)

Dromaius novaehollandiae 3.031 2.828 Buttemer et al. (1988)

Struthio camelus 3.627 3.139 Bertram and Burger (1981);

Swart et al. (1987); Swart and

Rahn (1988)

Crocodylia

Crocodylus acutus 2.135 1.963 Moore (1953); Lutz and

Dunbar-Cooper (1984)

168

Appendix 3.3. Museum specimens assessed for eggshell porosity in this study.

Institutional abbreviations: AIM, Auckland Institute and Museum, Auckland, New Zealand; CM,

Carnegie Museum of Natural History, Pittsburgh, Pennsylvania; HEC, Hirsch Egg Catalogue,

University of Colorado Museum, Boulder, Colorado; MCZ, Museum of Comparative Zoology,

Cambridge, Massachusetts; ROM, Royal Ontario Museum, Toronto, Canada; TMP, Royal Tyrrell

Museum of Palaeontology, Drumheller, Canada; UHR, Hokkaido University Museum, Sapporo,

Japan; YPM R., Herpetology Collection at the Yale Peabody Museum, New Haven, Connecticut;

ZEC, Zelenitsky Egg Catalogue, University of Calgary, Calgary, Canada.

Specimen Taxon Collection number

Modern eggshell Aix galericulata UHR 33131

Aix sponsa UHR 33128

Alectura lathami ZEC 137-1-3; ZEC 137-1-4; ZEC

137-1-5

Alligator mississippiensis YPM HERR. 015109; YPM HERR.

015110

Alligator sinensis YPM HERR. 018989

Ammoperdix heyi ROM 9516

Anas bahamensis ROM 9487

Anas discors ROM 7563

Anas platyrhynchos ROM 12821

Anhinga anhinga ROM 9438

Anser anser ZEC 221-1-2; ZEC 221-1-3

Branta canadensis ZEC 444-1-1; ZEC 444-1-2

169

Specimen Taxon Collection number

Bucephala islandica ROM 7739

Burhinus oedicnemus ROM 10983

Buteo rufinus ROM 8126

Caiman crocodilus UHR 33214

Caiman latirostris UHR 27367; YPM HERR. 017953;

YPM HERR. 018990

Caiman yacare UHR 33215; YPM HERR. 019030

Cairina moschata ZEC 290-1-1; ZEC 290-1-2

Chrysolophus amherstiae ROM 3659

Clangula hyemalis ROM 12427

Crocodylus moreletii YPM HERR. 018979; YPM HERR.

018980; YPM HERR. 018981

Crocodylus niloticus MCZ 26933; YPM HERR. 017955;

YPM HERR. 018982; YPM HERR.

018983; YPM HERR. 018984; YPM

HERR. 018985; YPM HERR. 018986;

ZEC 136 (HEC175)

Crocodylus porosus UHR 33210; YPM R17954; YPM

R17976

Crocodylus rhombifer YPM HERR. 011637; YPM HERR.

015102

Crocodylus siamensis YPM HERR. 018977; YPM HERR.

170

Specimen Taxon Collection number

018978

Egretta thula ROM 4772

Egretta tricolor ROM 4782

Eudocimus albus ROM 4818

Eudromia elegans ZEC 283-1-2

Falco naumanni ROM 5096

Falco tinnunculus ROM 10855

Fratercula arctica ROM 2354

Fulmarus glacialis ROM 3034

Gavialis gangeticus YPM HERR. 018824

Larus glaucescens ROM 5503

Larus heermanni ROM 8686

Larus ridibundus ROM 3780

Leipoa ocellata ZEC 218-2-1; ZEC218-2-2

Lophura nycthemera ROM 4622

Megapodius decollatus YPM 142019

Melanosushus niger CM41452 (ZEC303); MCZ 46554

Nycticorax nycticorax UHR 33121

Oceanodroma leucorhoa ROM 2796

Onychoprion fuscatus ROM 5632

Osteolaemus tetraspis YPM HERR. 018823; YPM HERR.

018988

171

Specimen Taxon Collection number

Paleosuchus palpebrosus CM41453 (ZEC304); YPM HERR.

017952

Paleosuchus trigonatus YPM HERR. 018987

Passer domesticus ROM 2377; ZEC 455-1-1; ZEC

455-1-2

Pavo cristatus UHR 33126

Phalacrocorax pelagicus ZEC 445-1-1; ZEC 445-1-2

Plegadis falcinellus ROM 8002

Pygoscelis adeliae ROM 11421

Rissa tridactyla ROM 356

Rynchops niger ROM 10003

Somateria m. mollissima ROM 10863

Spheniscus demersus ROM 9973

Sterna paradisaea ROM 5615

Sternula albifrons ROM 10864

Streptopelia turtur ROM 3665

Strix aluco ROM 3767

Syrmaticus soemmerringii UHR 33127

Tadorna tadorna ROM 9527

Thalasseus elegans ROM 13086

Thalasseus maximus ROM 5660

Tomistoma schlegelii YPM HERR. 018975; YPM HERR.

172

Specimen Taxon Collection number

018976

Tyto alba ROM 12631

Fossil/sub-fossil Euryapteryx sp. AIM LB6672; AIM LB6673 (Gill, eggshell 2000; Huynen et al., 2010)

Pachyornis geranoides AIM LB6675 (Gill, 2000; Huynen et

al., 2010)

Prismatoolithus levis/ TMP1994.179.1 (Holotype: Zelenitsky

Troodon formosus and Hills, 1996)

173

−1 Appendix 3.4. Eggshell porosity (Ap∙Ls ) and egg mass (M) of living and extinct archosaurs.

*Eggshell thickness was averaged between unincubated and hatched eggshells (Booth and

Seymour, 1987; Booth and Thompson, 1991). †Single pore area was calculated from pore radius reported by Lill (1987), assuming that cross-sectional pore canal is round. Egg length and breadth were taken from Schonwetter (1960–1967).

−1 Taxon log Ap∙Ls log M References

Accipitridae

Buteo rufinus 0.012 1.783 Ar and Rahn (1985);

this study

Anseriformes

Aix galericulata 0.590 1.547 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979); this study

Aix sponsa 0.711 1.642 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979); this study

Anas bahamensis 0.310 1.543 Schonwetter

(1960–1967); Hoyt

et al. (1979); this

study

Anas discors -0.259 1.405 Schonwetter

174

−1 Taxon log Ap∙Ls log M References

(1960–1967); Hoyt

et al. (1979); this

study

Anas fulvigula 1.051 1.747 Hoyt et al. (1979)

Anas gracilis 0.527 1.515 Hoyt et al. (1979)

Anas platyrhynchos 0.537 1.895 Schonwetter

(1960–1967);

Rokitka and Rahn

(1987); Balkan et al.

(2006); this study

Anser anser 0.857 2.192 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979); Rokitka and

Rahn (1987); this

study

Anser brachyrhynchus 0.817 2.144 Hoyt et al. (1979)

Anser cygnoides 1.264 2.166 Hoyt et al. (1979)

Anser erythropus 0.902 2.090 Hoyt et al. (1979)

Anser fabalis 0.813 2.183 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

175

−1 Taxon log Ap∙Ls log M References

(1979)

Branta canadensis 1.138 2.167 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985); this study

Branta h. minima 0.383 2.002 Hoyt et al. (1979)

Branta leucopsis 0.651 2.028 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979)

Branta sandvicensis 0.905 2.189 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979)

Bucephala islandica 0.822 1.827 Hoyt et al. (1979);

this study

Cairina moschata 0.742 1.898 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985); this study

Cereopsis novaehollandiae 0.368 2.127 Schonwetter

(1960–1967); Tullett

176

−1 Taxon log Ap∙Ls log M References

(1976); Ar and Rahn

(1985)

Chloephaga melanoptera 0.728 2.049 Carey et al. (1990)

Clangula hyemalis 0.284 1.661 Rahn et al. (1983);

this study

Cyanochen cyanoptera 0.663 1.918 Hoyt et al. (1979)

Dendrocygna arborea 0.703 1.777 Hoyt et al. (1979)

Dendrocygna autumnalis 0.352 1.631 Hoyt et al. (1979)

Dendrocygna bicolor 0.587 1.736 Hoyt et al. (1979)

Lophodytes cucullatus 0.265 1.718 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979)

Mergus merganser 0.848 1.841 Hoyt et al. (1979)

Mergus serrator 0.713 1.823 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979)

Somateria mollissima 0.905 2.033 Rahn et al. (1983);

this study

Tadorna cana 0.898 1.919 Schonwetter

(1960–1967); Tullett

177

−1 Taxon log Ap∙Ls log M References

(1976); Ar and Rahn

(1985)

Tadorna tadorna 0.888 1.903 Hoyt et al. (1979);

this study

Tadorna variegata 1.095 1.952 Schonwetter

(1960–1967); Tullett

(1976); Hoyt et al.

(1979)

Charadriiformes

Alca torda 0.277 1.968 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Brachyramphus marmoratus 0.433 1.585 Nelson (1997);

Zimmermann and

Hipfner (2007)

Burhinus oedicnemus 0.677 1.525 Ar and Rahn (1985);

this study

Cepphus columba 0.589 1.756 Zimmermann and

Hipfner (2007)

Cerorhinca monocerata 0.317 1.899 Hipfner et al. (2004);

Zimmermann and

178

−1 Taxon log Ap∙Ls log M References

Hipfner (2007)

Fratercula arctica 0.683 1.778 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985); this study

Fratercula cirrhata 0.703 1.954 Tullett (1976);

Zimmermann and

Hipfner (2007)

Haematopus ostralegus 0.256 1.618 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Larus argentatus 0.695 1.940 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Larus canus 0.593 1.699 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Larus fuscus 0.976 1.924 Schonwetter

(1960–1967); Tullett

179

−1 Taxon log Ap∙Ls log M References

(1976); Ar and Rahn

(1985)

Larus glaucescens 0.894 1.991 Ar and Rahn (1985);

this study

Larus heermanni 0.610 1.730 Rahn and Dawson

(1979); Ar and Rahn

(1985); this study

Larus marinus 0.878 2.057 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Larus ridibundus 0.300 1.555 Ar and Rahn (1985);

this study

Numenius phaeopus 0.774 1.728 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Onychoprion fuscatus 0.408 1.531 Ar and Rahn (1985);

this study

Pluvialis apricaria 0.294 1.513 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

180

−1 Taxon log Ap∙Ls log M References

(1985)

Ptychoramphus aleuticus -0.138 1.465 Zimmermann and

Hipfner (2007)

Rissa tridactyla 0.393 1.699 Ar and Rahn (1985);

this study

Rynchops niger 0.043 1.422 Grant et al. (1984);

Ar and Rahn (1985);

this study

Stercorarius skua 0.767 1.980 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Sterna paradisaea -0.319 1.255 Ar and Rahn (1985);

this study

Sternula albifrons -0.396 0.968 Ar and Rahn (1985);

this study

Synthliboramphus antiquus 0.060 1.651 Gaston (1994);

Zimmermann and

Hipfner (2007)

Thalasseus elegans 0.297 1.613 Ar and Rahn (1985);

this study

Thalasseus maximus 0.769 1.789 Schonwetter

181

−1 Taxon log Ap∙Ls log M References

(1960–1967);

Rokitka and Rahn

(1987); this study

Uria aalge 0.749 2.048 Tullett (1976);

Zimmermann and

Hipfner (2007)

Ciconiiformes

Egretta thula 0.333 1.354 Ar and Rahn (1985);

this study

Egretta tricolor 0.092 1.407 Ar and Rahn (1985);

this study

Eudocimus albus 0.213 1.696 Ar and Rahn (1985);

this study

Nycticorax nycticorax 0.147 1.618 Ar and Rahn (1985);

this study

Plegadis falcinellus 0.424 1.571 Ar and Rahn (1985);

this study

Columbiformes

Columba livia 0.050 1.193 Tullett (1976); Arad

et al. (1988)

Streptopelia turtur -0.479 0.919 Ar and Rahn (1985);

this study

182

−1 Taxon log Ap∙Ls log M References

Falconidae

Falco naumanni -0.206 1.035 Ar and Rahn (1985);

this study

Falco tinnunculus -0.321 1.258 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1978); this study

Galliformes

Alectura lathami 1.515 2.299 Ar and Rahn (1985);

Booth and

Thompson (1991);

this study

Ammoperdix heyi -0.652 1.154 Ar and Rahn (1985);

this study

Chrysolophus amherstiae 0.724 1.484 Ar and Rahn (1985);

this study

Coturnix coturnix 0.160 0.982 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Gallus gallus 0.273 1.752 Schonwetter

(1960–1967); Tullett

183

−1 Taxon log Ap∙Ls log M References

(1976); Rokitka and

Rahn (1987)

Leipoa ocellata 1.673 2.197 Schonwetter

(1960–1967); Tullett

(1976); Booth

(1987); Booth and

Seymour (1987); this

study

Lophophorus impejanus 0.337 1.816 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Lophura nycthemera 0.877 1.601 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985); this study

Megapodius decollatus 1.428 2.069 Jones et al. (1995);

this study

Meleagris gallopavo 0.741 1.863 Schonwetter

(1960–1967); Tullett

(1976); Rokitka and

Rahn (1987)

184

−1 Taxon log Ap∙Ls log M References

Numida meleagris 0.235 1.689 Tullett (1976); Ancel

and Girard (1992)

Pavo cristatus 1.163 1.978 Ar and Rahn (1985);

this study

Phasianus colchicus 0.357 1.498 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Syrmaticus soemmerringii 0.436 1.498 Ar and Rahn (1985);

this study

Passeriformes

Agelaius phoeniceus -1.358 0.633 Rahn et al. (1977b);

Jaeckle et al. (2012)

Menura novaehollandiae 0.044 1.789 Lill (1987); Higgins

et al. (2001)

Molothrus ater -1.154 0.522 Ar et al. (1974);

Jaeckle et al. (2012)

Passer domesticus -0.867 0.441 Tullett (1976); Ar

and Rahn (1985);

this study

Spiza americana -1.682 0.450 Temple (2002);

Jaeckle et al. (2012)

185

−1 Taxon log Ap∙Ls log M References

Turdus merula -0.278 0.828 Tullett (1976); Ar

and Rahn (1985);

Higgins et al. (2006)

Pelecaniformes

Anhinga anhinga 0.469 1.560 Colacino et al.

(1985); this study

Morus bassanus 0.397 2.025 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Phalacrocorax auritus 0.450 1.696 Tullett (1976); Ar

and Rahn (1985)

Phalacrocorax carbo 0.392 1.763 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Phalacrocorax pelagicus 0.444 1.598 Ar and Rahn (1985);

this study

Procellariiformes

Diomedea exulans 1.357 2.699 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

186

−1 Taxon log Ap∙Ls log M References

(1985)

Fulmarus glacialis 1.129 2.004 Rahn et al. (1984);

this study

Oceanodroma leucorhoa -0.295 1.021 Ar and Rahn (1985);

this study

Puffinus pacificus 0.844 1.778 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Puffinus puffinus 0.591 1.766 Tullett (1976); Lee

and Haney (1996)

Puffinus tenuirostris 0.927 1.942 Schonwetter

(1960–1967);

Fitzherbert (1985)

Sphenisciformes

Aptenodytes forsteri 1.188 2.653 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985)

Aptenodytes patagonicus 1.149 2.481 Tullett (1976);

Handrich (1989)

Eudyptes robustus 0.827 2.065 Massaro and Davis

187

−1 Taxon log Ap∙Ls log M References

(2005)

Pygoscelis adeliae 0.899 2.096 Rahn and Hammel

(1982); this study

Spheniscus demersus 0.692 2.003 Schonwetter

(1960–1967); Tullett

(1976); Ar and Rahn

(1985); this study

Strigiformes

Strix aluco 0.207 1.558 Ar and Rahn (1985);

this study

Tyto alba 0.126 1.272 Ar and Rahn (1985);

this study

Struthioniformes

Apteryx australis 1.024 2.544 Calder III et al.

(1978);

Silyn-Roberts (1983)

Tinamiformes

Eudromia elegans 0.408 1.554 Ar and Rahn (1985);

this study

Crocodylia

Alligator mississippiensis 1.497 1.859 Packard et al. (1979);

Wink et al. (1990);

188

−1 Taxon log Ap∙Ls log M References

Marzola et al.

(2015); this study

Alligator sinensis 1.121 1.713 Ferguson (1985);

Wink and Elsey

(1994); this study

Caiman crocodilus 1.175 1.771 Ferguson (1985);

this study

Caiman latirostris 1.736 1.829 Stoker et al. (2013);

this study

Caiman yacare 1.264 1.877 Ferguson (1985);

this study

Crocodylus mindorensis 1.462 1.867 Marzola et al.

(2015);

Thorbjarnarson

(1996)

Crocodylus moreletii 1.130 1.839 Platt et al. (2008);

this study

Crocodylus niloticus 1.300 2.041 Ferguson (1985);

this study

Crocodylus porosus 1.325 2.053 Ferguson (1985);

this study

Crocodylus rhombifer 1.325 2.018 Ferguson (1985);

189

−1 Taxon log Ap∙Ls log M References

Thorbjarnarson

(1996); this study

Crocodylus siamensis 1.245 2.029 Ferguson (1985);

Thorbjarnarson

(1996); this study

Gavialis gangeticus 2.080 2.208 Ferguson (1985);

Thorbjarnarson

(1996); this study

Melanosushus niger 1.548 2.157 Ferguson (1985);

Thorbjarnarson

(1996); this study

Osteolaemus tetraspis 0.766 1.715 Ferguson (1985);

this study

Paleosuchus palpebrosus 1.522 1.837 Ferguson (1985);

Marzola et al.

(2015); this study

Paleosuchus trigonatus 1.861 1.858 Rivas et al. (2001);

this study

Tomistoma schlegelii 1.658 2.146 Ferguson (1985);

Thorbjarnarson

(1996); this study

190

Appendix 3.5. Nest type classification for living archosaur species.

Taxon Nest type References

Accipitridae Buteo rufinus Open del Hoyo et al. (1994);

Hayman and Hume

(2007)

Anseriformes Aix galericulata Open Harrison (1975); Kear

(2005)

Aix sponsa Open Delacour (1959); Hepp

and Bellrose (1995);

Baicich and Harrison

(1997); Kear (2005)

Anas bahamensis Open Kear (2005)

Anas discors Open Baicich and Harrison

(1997)

Anas fulvigula Open Stieglit and Wilson

(1968); Bellrose (1976)

Anas gracilis Open Serventy and Whittell

(1962); Pizzey (1980)

Anas platyrhynchos Open Kear (2005)

Anser anser Open Harrison (1975); Cramp

(1977)

Anser brachyrhynchus Open Kear (2005)

Anser cygnoides Open Kear (2005)

191

Taxon Nest type References

Anser erythropus Open Harrison (1975)

Anser fabalis Open Harrison (1975)

Branta canadensis Open Harrison (1975)

Branta h. minima Open Harrison (1975)

Branta leucopsis Open Harrison (1975)

Branta sandvicensis Open Merne (1974); Banko et

al. (1999)

Bucephala islandica Open Harrison (1975)

Cairina moschata Open Baicich and Harrison

(1997)

Cereopsis novaehollandiae Open Pizzey (1980); Kear

(2005)

Chloephaga melanoptera Open Carey et al. (1990); Kear

(2005)

Clangula hyemalis Open Harrison (1975)

Cyanochen cyanoptera Open del Hoyo et al. (1992);

Kear (2005)

Dendrocygna arborea Open Kear (2005)

Dendrocygna autumnalis Open Baicich and Harrison

(1997)

Dendrocygna bicolor Open Baicich and Harrison

(1997)

192

Taxon Nest type References

Lophodytes cucullatus Open Bellrose (1976); del

Hoyo et al. (1992)

Mergus merganser Open Harrison (1975)

Mergus serrator Open Harrison (1975)

Somateria mollissima Open Harrison (1975);

Mehlum (1991)

Tadorna cana Open Kear (2005)

Tadorna tadorna Open Harrison (1975)

Tadorna variegata Open Williams (1979); del

Hoyo et al. (1992)

Charadriiformes Alca torda Open Baicich and Harrison

(1997)

Brachyramphus marmoratus Open Baicich and Harrison

(1997)

Burhinus oedicnemus Open Harrison (1975)

Cepphus columba Open Baicich and Harrison

(1997)

Cerorhinca monocerata Open Baicich and Harrison

(1997)

Fratercula arctica Open Harrison (1975)

Fratercula cirrhata Open Baicich and Harrison

(1997)

193

Taxon Nest type References

Haematopus ostralegus Open Harrison (1975)

Larus argentatus Open Harrison (1975)

Larus canus Open Harrison (1975); Burger

and Gochfeld (1987)

Larus fuscus Open Harrison (1975)

Larus glaucescens Open Baicich and Harrison

(1997)

Larus heermanni Open Baicich and Harrison

(1997)

Larus marinus Open Harrison (1975)

Larus ridibundus Open Harrison (1975); Cramp

(1983)

Numenius phaeopus Open Harrison (1975)

Onychoprion fuscatus Open Baicich and Harrison

(1997)

Pluvialis apricaria Open Harrison (1975)

Ptychoramphus aleuticus Open Baicich and Harrison

(1997)

Rissa tridactyla Open Harrison (1975); Baicich

and Harrison (1997)

Rynchops niger Open Baicich and Harrison

(1997)

194

Taxon Nest type References

Stercorarius skua Open Harrison (1975)

Sterna paradisaea Open Harrison (1975); Cramp

(1985)

Sternula albifrons Open Harrison (1975)

Synthliboramphus antiquus Open Baicich and Harrison

(1997)

Thalasseus elegans Open Baicich and Harrison

(1997)

Thalasseus maximus Open Baicich and Harrison

(1997)

Uria aalge Open Harrison (1975)

Ciconiiformes Egretta thula Open Baicich and Harrison

(1997)

Egretta tricolor Open Baicich and Harrison

(1997)

Eudocimus albus Open Baicich and Harrison

(1997)

Nycticorax nycticorax Open Harrison (1975)

Plegadis falcinellus Open Harrison (1975)

Columbiformes Columba livia Open Harrison (1975)

Streptopelia turtur Open Harrison (1975)

Falconidae Falco naumanni Open Harrison (1975); del

195

Taxon Nest type References

Hoyo et al. (1994); Snow

and Perrins (1998)

Falco tinnunculus Open Harrison (1975); Snow

and Perrins (1998)

Galliformes Alectura lathami Covered Pizzey (1980)

Ammoperdix heyi Open Harrison (1975); Snow

and Perrins (1998)

Chrysolophus amherstiae Open Beebe (1931); del Hoyo

et al. (1994)

Coturnix coturnix Open Harrison (1975)

Gallus gallus Open Pizzey (1980)

Leipoa ocellata Covered Pizzey (1980)

Lophophorus impejanus Open Beebe (1931)

Lophura nycthemera Open Harrison (1975)

Megapodius decollatus Covered Jones et al. (1995)

Meleagris gallopavo Open Harrison (1975)

Numida meleagris Open Baicich and Harrison

(1997)

Pavo cristatus Open del Hoyo et al. (1994);

Whistler (1949)

Phasianus colchicus Open Harrison (1975)

Syrmaticus soemmerringii Open Yamashina (1961)

196

Taxon Nest type References

Passeriformes Agelaius phoeniceus Open Baicich and Harrison

(1997)

Menura novaehollandiae Open Lill (1979); Pizzey

(1980)

Molothrus ater Open Baicich and Harrison

(1997)

Passer domesticus Open Harrison (1975)

Spiza americana Open del Hoyo et al. (2011)

Turdus merula Open Harrison (1975)

Pelecaniformes Anhinga anhinga Open Harrison (1975)

Morus bassanus Open Baicich and Harrison

(1997)

Phalacrocorax auritus Open Baicich and Harrison

(1997)

Phalacrocorax carbo Open Baicich and Harrison

(1997)

Phalacrocorax pelagicus Open Baicich and Harrison

(1997)

Procellariiformes Diomedea exulans Open del Hoyo et al. (1992)

Fulmarus glacialis Open Harrison (1975); Cramp

(1977)

Oceanodroma leucorhoa Open Harrison (1975)

197

Taxon Nest type References

Puffinus pacificus Open Howell and

Bartholomew (1961);

Whittow (1997)

Puffinus puffinus Open Baicich and Harrison

(1997)

Puffinus tenuirostris Open Bradley et al. (2000)

Sphenisciformes Aptenodytes forsteri Open del Hoyo et al. (1992)

Aptenodytes patagonicus Open Handrich (1989)

Eudyptes robustus Open Warham (1974)

Pygoscelis adeliae Open Goodfellow (1977)

Spheniscus demersus Open Seddon and van Heezik

(1991); Kemper et al.

(2007)

Strigiformes Strix aluco Open Harrison (1975)

Tyto alba Open Harrison (1975)

Struthioniformes Apteryx australis Open Calder III (1979);

Colbourne (2002)

Tinamiformes Eudromia elegans Open Mezquida (2001); Davies

(2002)

Crocodylia Alligator mississippiensis Covered Brazaitis and Watanabe

(2011)

Alligator sinensis Covered Brazaitis and Watanabe

198

Taxon Nest type References

(2011)

Caiman crocodilus Covered Brazaitis and Watanabe

(2011)

Caiman latirostris Covered Brazaitis and Watanabe

(2011)

Caiman yacare Covered Brazaitis and Watanabe

(2011)

Crocodylus mindorensis Covered Brazaitis and Watanabe

(2011)

Crocodylus moreletii Covered Brazaitis and Watanabe

(2011)

Crocodylus niloticus Covered Brazaitis and Watanabe

(2011)

Crocodylus porosus Covered Brazaitis and Watanabe

(2011)

Crocodylus rhombifer Covered Brazaitis and Watanabe

(2011)

Crocodylus siamensis Covered Brazaitis and Watanabe

(2011)

Gavialis gangeticus Covered Brazaitis and Watanabe

(2011)

Melanosushus niger Covered Brazaitis and Watanabe

199

Taxon Nest type References

(2011)

Osteolaemus tetraspis Covered Brazaitis and Watanabe

(2011)

Paleosuchus palpebrosus Covered Brazaitis and Watanabe

(2011)

Paleosuchus trigonatus Covered Brazaitis and Watanabe

(2011)

Tomistoma schlegelii Covered Brazaitis and Watanabe

(2011)

200

Appendix 3.6. Cladogram of 196 living bird species for comparisons of MH2O

(phylogenetically-generalized least-squares model in Appendix 3.7).

Modified from Sheldon (1987); Livezey (1995, 1996); Sheldon and Slikas (1997); Kennedy and

Spencer (2000, 2004); Sheldon et al. (2000); Wink and Heidrich (2000); Johnson et al. (2001);

Livezey (2001); Kennedy and Page (2002); Fjeldsa (2004); Krebs and Putland (2004); Thomas et al. (2004); Bertelli and Giannini (2005); de Kloet and de Kloet (2005); Lerner and Mindell

(2005); Crowe et al. (2006); Ericson et al. (2006); Jonsson and Fjeldsa (2006); Ksepka et al.

(2006); Griffiths et al. (2007); Proudfoot et al. (2007); Wright et al. (2008); Gonzalez et al.

(2009); Mayr (2010); McCracken et al. (2010); Smith (2010); Yang et al. (2010); Jarvis et al.

(2014).

201

202

Appendix 3.7. Bivariate plot of daily loss of water vapor and egg mass in living species.

Ordinary least-squares (OLS) and phylogenetically-generalized least-squares (PGLS) regressions as well as the 95% confidence intervals (CIs) were conducted for living archosaur species using

IBM SPSS Statistics v. 22.0.0 (IBM SPSS Inc.) and PDAP module v.1.15 (Midford et al., 2010) of the software Mesquite 3.02 (Maddison and Maddison, 2010). For the phylogenetic approach, a phylogenetic tree of 196 species was reconstructed (Appendix 3.6) and tree lengths were assigned based on Pagel's method (Pagel, 1992) and divergence time. White and black dots represent open

2 and covered nesters, respectively. Note that log MH2O is strongly correlated with log M (r = 0.888 for OLS, 0.638 for PGLS with Pagel's branch length method, and 0.670 for PGLS with the divergence time method).

203

Appendix 3.8. Cladogram of 127 living archosaur species for pcANCOVA.

Modified from Sheldon (1987); Livezey (1995, 1996); Sheldon and Slikas (1997); Kennedy and

Spencer (2000); Sheldon et al. (2000); Wink and Heidrich (2000); Johnson et al. (2001); Livezey

(2001); Kennedy and Page (2002); Fjeldsa (2004); Thomas et al. (2004); Bertelli and Giannini

(2005); de Kloet and de Kloet (2005); Lerner and Mindell (2005); Crowe et al. (2006); Ericson et al. (2006); Jonsson and Fjeldsa (2006); Ksepka et al. (2006); Griffiths et al. (2007); Wright et al.

(2008); Gonzalez et al. (2009); Mayr (2010); McCracken et al. (2010); Smith (2010); Yang et al.

(2010); Oaks (2011); Jarvis et al. (2014); Mitchell et al. (2014).

204

205

Appendix 3.9. Cladogram of 133 living and extinct archosaur taxa for pFDA.

Modified from Sheldon (1987); Livezey (1995, 1996); Sheldon and Slikas (1997); Kennedy and

Spencer (2000); Sheldon et al. (2000); Wink and Heidrich (2000); Johnson et al. (2001); Livezey

(2001); Kennedy and Page (2002); Fjeldsa (2004); Thomas et al. (2004); Bertelli and Giannini

(2005); de Kloet and de Kloet (2005); Lerner and Mindell (2005); Crowe et al. (2006); Ericson et al. (2006); Jonsson and Fjeldsa (2006); Ksepka et al. (2006); Griffiths et al. (2007); Wright et al.

(2008); Gonzalez et al. (2009); Mayr (2010); McCracken et al. (2010); Smith (2010); Yang et al.

(2010); Nesbitt (2011); Oaks (2011); Carrano et al. (2012); Turner et al. (2012); Jarvis et al.

(2014); Mitchell et al. (2014).

206

207

Appendix 4.1. Mean nest temperature (Tnest, °C), mean ambient air temperature (Tair, °C), and the difference between Tnest and

Tair (°C) of living covered nesters. Note that * indicates mean values were calculated in this study using the data provided by the papers.

Incubation heat source Taxon Tnest Tair Difference References Microbial respiration Aepypodius arfakianus 31.100 21.800 9.300 Sinclair (2001) Aepypodius sp. 37.800 15.600 22.200 Frith (1956) Alectura lathami 33.300 15.100 18.200 Jones (1988) 32.950 17.800 15.150 Seymour and Bradford (1992) Leipoa ocellata 32.626* 27.185* 5.441 Lewis (1940) Megapodius decollatus 31.100 22.100 9.000 Sinclair (2001) Megapodius freycinet 34.400 26.750 7.650 Lincoln (1974) 33.500* 26.750* 6.750 Stuebing and Johan (1986) Megapodius laperouse 31.300 28.500 2.800 Wiles and Conry (2001) Megapodius nicobariensis 32.440 28.170 4.270 Sivakumar and Sankaran (2003) Talegalla jobiensis 30.600 21.800 8.800 Sinclair (2001) Alligator mississippiensis 34.034 28.020 6.014 McIlhenny (1934) 28.200 30.700 -2.500 Joanen (1969) 29.833 28.267 1.566 Chabreck (1973) Caiman c. crocodilus 29.900* 29.571* 0.329 Staton and Dixon (1977) Crocodylus cataphractus 30.700* 25.500* 5.200 Waitkuwait (1985) Melanosuchus niger 31.200 27.900 3.300 Villamarln-Jurado and Suarez (2007) Geothermal heat Macrocephalon maleo 35.000* 28.250* 6.750 MacKinnon (1978)

208

Incubation heat source Taxon Tnest Tair Difference References Megapodius eremita 33.000 32.000 1.000 Roper (1983)

Megapodius pritchardii 32.500* 27.000* 5.500 Goth and Vogel (1996–1997) Solar radiation Eulipoa wallacei (dry 32.600 27.442* 5.158 Heij et al. (1997) season values) Eulipoa wallacei (rainy 32.500 26.285* 6.215 Heij et al. (1997) season values) Macrocephalon maleo 36.000 31.000 5.000 MacKinnon (1978) Megapodius laperouse 33.500* 30.2500* 3.250 Glass et al. (1988) Crocodylus acutus 31.670 30.689 0.981 Charruau (2012) Crocodylus niloticus 32.725* 31.213* 1.512 Modha (1967)

209

Appendix 4.2. Nest type, nest material, and heat source of incubation in living covered nesters. Abbreviations: C, clay; H, in-filled hole nest; I, inorganic heat source; M, mound nest; O, organic heat source (with possibly some inroganic heat); PS, plant material and/or soil; Sa, sand. Order Species Nest Nest Heat References type material source Galliformes Aepypodius arfakianus M PS O Sinclair (2001) Aepypodius bruijni M PS O Mauro (2005) Alectura lathami M PS O Le Souef (1899); Frith (1956) Eulipoa wallacei H Sa I Ripley (1960); Rand and Gilliard (1967); Dekker et al. (1995); Heij et al. (1997); Baker (1999); Baker and Dekker (2000) Leipoa ocellata M Sa O Le Souef (1899); Frith (1956, 1959) Macrocephalon maleo H Sa I MacKinnon (1978); Dekker (1988); Argeloo and Dekker (1996) H PS I Dekker (1988) Megapodius bernsteinii M PS I Indrawan et al. (1998) M Sa I Indrawan et al. (1998) Megapodius cumingii M Sa O Jones et al. (1995) Megapodius decollatus M PS O Sinclair (2001) Megapodius eremita H Sa I Roper (1983) Megapodius forstenii M PS Jones et al. (1995) M Sa Jones et al. (1995) M PS Stresemann (1914)

210

Order Species Nest Nest Heat References type material source Megapodius freycinet H Sa O MacKinnon (1978) M Sa I Lincoln (1974) M PS Ripley (1960) M PS O Lincoln (1974) Megapodius laperous H Sa Falanruw (1975) laperous Megapodius l. senex H PS I Glass et al. (1988) M PS Stinson and Glass (1992) M Sa O Wiles and Conry (2001) Megapodius layardi H PS O Bowen (1996) H PS I Bregulla (1992) Megapodius M Sa O Sankaran (1995); Sivakumar and Sankaran (2003) nicobariensis Megapodius pritchardii H Sa I Friedlander (1899) H PS I Kellers (1931); Weir (1973) Megapodius reinwardt M PS O Le Souef (1899); Crome and Brown (1979) M Sa O Crome and Brown (1979) Talegalla cuvieri M PS Gilliard and LeCroy (1970) Talegalla fuscirostris M PS Jones et al. (1995) Talegalla jobiensis M PS O Sinclair (2001)

211

Order Species Nest Nest Heat References type material source Crocodylia Alligator M PS O Joanen (1969); Chabreck (1975); Metzen (1977); Goodwin mississippiensis and Marion (1978) Alligator sinensis M PS O Thorbjarnarson et al. (2001) Caiman crocodilus M PS O Ouboter and Nanhoe (1987); Allsteadt (1994) M PS O Staton and Dixon (1977) Caiman latirostris M PS Montini et al. (2006) Caiman yacare M PS Crawshaw and Schaller (1980); Cintra (1988) Crocodylus acutus H Sa I Platt and Thorbjarnarson (2000); Charruau (2012) M PS Ogden (1978); Mazzotti (1989) M Sa Ogden (1978); Lutz and Dunbar-Vooper (1984); Kushlan and Mazzotti (1989); Mazzotti (1989) Crocodylus M PS O Waitkuwait (1985) cataphractus Crocodylus intermedius H Sa Medem (1981); Bolhm (1982); Thorbjarnarson and Hernandez (1993); Seijas (1998); Espinosa-Blanco et al. (2013) M PS Medem (1981) Crocodylus johnstoni H Sa I Webb et al. (1983); Somaweera and Shine (2013) Crocodylus mindorensis M PS van Weerd et al. (2006); van Weerd (2010) Crocodylus moreletii M PS O Platt et al. (2008); Lopez-Luna et al. (2015)

212

Order Species Nest Nest Heat References type material source Crocodylus niloticus H C Pooley (1969) H Sa I Cott (1961); Modha (1967); Pooley (1969); Kofron (1989); Swanepoel et al. (2000) Crocodylus M PS O Neil (1946); Hall and Johnson (1987) novaeguineae Crocodylus palustris H C Whitaker and Whitaker (1976) H Sa Whitaker and Whitaker (1976); Rao and Gurjwar (2013) Crocodylus porosus H PS Whitaker and Whitaker (1976) M PS O Webb et al. (1977, 1983) Crocodylus siamensis M PS Bezuijen et al. (2013) Gavialis gangeticus H Sa (1980); Rao (1988); Nair and Katdare (2013); Rao et al. (2013); Rao and Gurjwar (2013) Melanosuchus niger M PS O Herron et al. (1990); Villamarln-Jurado and Suarez (2007) Osteolaemus tetraspis M PS Riley and Huchzermeyer (1999) Paleosuchus M PS Medem (1958, 1971) palpebrosus Paleosuchus trigonatus M PS O Magnusson et al. (1985); Rivas et al. (2001) Tomistoma schlegelii M PS Staniewicz (2011)

213

Appendix 4.3. Lithology of in-situ dinosaur eggs (clutches and nesting grounds). *Indicates pedogenetic feature(s) is identified. Abbreviation: NA, not available; TMP, Royal Tyrrell Museum of Palaeontology, Drumheller, Canada. Ootaxon/taxon Material Lithology Formation Locality References Cairanoolithidae Cairanoolithus dughii Clutch(es) Fine NA Commune de Rousset, Williams et al. (1984) near Cairanne, Aix-en-Provence, France Cairanoolithus cf. Clutch(es) Fine Coll de Nargo, Lleida, Sellés et al. (2013) roussetensis , Spain Dendroolithidae (/Phaceloolithidae) D. cf. dendriticus Clutch(es) Fine Chichengshan Shuangtang Tiantai Fang et al. (1998, 2000) Basin, Zhejiang, China D. fengguangcunensis Clutch(es) Coarse Dongyuan Heyuan Basin, Fang et al. (2005) , China D. sp. Clutch(es) Coarse Nemegt Khaichin Ula, Gobi, Suzuki and Watabe Mongolia (2000b) Phaceloolithus Clutch(es) Fine Fenshuiao Western Dongting Zeng and Zhang (1979) hunanensis Basin, Hunan, China Dendroolithidae indet. Clutch(es) Coarse ?Djadokhta Toosgot area, Abdrant Watabe (2004)

214

Ootaxon/taxon Material Lithology Formation Locality References Nuru, Gobi, Mongolia Dendroolithidae indet. Clutch(es) Coarse Shiwa Shiwa Bay, Whaseong, Lee (2003) Gyeonggi, South Korea Dendroolithidae indet. Clutch(es) Fine Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 8-1G) South Korea Dendroolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Locations 8-2 and South Korea 8-3) Dendroolithidae indet. Clutch(es) Fine Baynshire Bayn Shire, Gobi, Watabe et al. (2010) Mongolia Torvosaurus Clutch(es) with Fine Lourinha Walen, Porto das Araujo et al. (2013) embryonic bones Barcas, Portugal Dictyoolithidae Dictyoolithus Clutch(es) Fine Quantou Gongzhuling area, Jilin, Wang et al. (2006) gongzhulingensis China D. hongpoensis Clutch(es) Coarse Chichengshan Lishui Basin, Zhejiang, Jin et al. (2010) China Protodictyoolithus Clutch(es) Fine Goseong Yeondo, Dosanmyeon, Kim et al. (2011) neixiangensis Tongyeong City, Gyeongsang, South Korea

215

Ootaxon/taxon Material Lithology Formation Locality References P. neixiangensis Clutch(es) Fine Goseong Southwest coast Kim et al. (2011) Pyeongri, Dosanmyeon, Tongyeong City, Gyeongsang, South Korea Elongatoolithidae Ellipsoolithus Clutch(es) Coarse* Lameta Lavariya Muwada, Mohabey (1998) khedaensis Kheda, Gujarat, India Elongatoolithus Clutch(es) Fine Yuanpu and Tangmienling, Shuikou, Young (1965); Zhao andrewsi Pingling Nanxiong Basin, (1975) Guangdong, China E. elongatus Clutch(es) Fine Yuanpu and Nanxiong Basin, Young (1965); Zhao Pingling Guangdong, China (1975) E. elongatus Clutch(es) Coarse NA Baiyuan, Pingxiang, Wu and Peng (2003) , China E. magnus Clutch(es) Fine Fenshuiao Western Dongting Zeng and Zhang (1979) Basin, Hunan, China Macroelongatoolithus A pair of eggs Fine Wayan Bonneville County, Krumenacker et al. carlylei , USA (2016) M. goseongensis Clutch(es) Fine Goseong Ddabakseom, Pyeongri, Kim et al. (2011) Dosanmyeon,

216

Ootaxon/taxon Material Lithology Formation Locality References Tongyeong City, Gyeongsang, South Korea M. xixiaensis Clutch(es) Fine Gaogou Xixia Basin, , Wang and Zhou (1995) China M. xixiaensis Clutch(es) Fine NA Zhouzhi Hill, Zhaoying Li et al. (1995) Village, Yangcheng Town, Xixia Basin, Henan, China M. xixiaensis Clutch(es) Fine Chichengshan Shuangtang, Tiantai Wang et al. (2010) Basin, Zhejiang, China M. xixiaensis Clutch(es) Fine Gyeongsang Located on the island of Huh et al. (2014) Basin Aphae-do in Shinan-gun, Jeollanam-do, southwestern Korea Macroolithus rugustus Clutch(es) Fine Yuanpu and Lashuyuan, Wuching, Young (1965); Zhao Pingling Nanxiong Basin, (1975) Guangdong, China M. rugustus? Clutch(es) Coarse* Nemegt Khaychin-Ula, Gobi, Sochava (1969); Mongolia Mikhailov (1994b)

217

Ootaxon/taxon Material Lithology Formation Locality References M. yaotunensis Clutch(es) Fine Yuanpu and Yaotun, Nanxiong Young (1965); Zhao Pingling Basin, Guangdong, (1975) China M. yaotunensis Clutch(es) Fine Yuanpu and Yaotun, Nanxiong Young (1965); Zhao Pingling Basin, Guangdong, (1975) China Nanhsiungoolithus Clutch(es) Fine Yuanpu and Nanxiong, Nanxiong Young (1965); Zhao chuetienensis Pingling Basin, Guangdong, (1975) China Trachoolithus sp. Clutch(es) Coarse* Lameta Lavariay Muwada, Mohabey (2000, 2001) Gujarat, India Elongatoolithidae indet. Clutch(es) Coarse Shiwa Shiwa Bay, Whaseong, Lee (2003) Gyeonggi, South Korea Elongatoolithidae indet. Clutch(es) Coarse Djadokhta Bayn Dzak, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae indet. Clutch(es) Coarse Baynshire Bayn Shire, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae indet. Clutch(es) Coarse NA Tel Ulan Chaltsai Watabe et al. (2010) (Mogoin Daatsyn Khuduk), Mongolia E. frustrabilis Clutch(es) Coarse Djadokhta Shabarakh Usu MNG, Brown and Schlaikjer

218

Ootaxon/taxon Material Lithology Formation Locality References Mongolia (1940); van Straelen (1925) E. frustrabilis? Clutch(es) with Coarse Djadokhta Xanadu, Ukhaa Tolgod, Norell et al. (1994, 2001) embryonic bones south-central Mongolia osmolskae Clutch(es) with Coarse Djadokhta Ankylosaur Flats, Ukhaa Clark et al. (1999); adult skeleton Tolgod, Mongolia Grellet-Tinner et al. (2006) barsboldi Clutch(es) with Coarse Baruungoyot Northern Sayr of the Fanti et al. (2012) adult skeleton Nemegt locality, Mongolia philoceratops Clutch(es) with Coarse Djadokhta North Canyon, Bayan Dong and Currie (1996) (cf. Machairasaurus) adult skeleton Mandahu, Inner Mongolia, China Elongatoolithidae indet. Clutch(es) with Fine Nemegt Bugin-Tsav, Gobi, Weishampel et al. (2008) embryonic bones Mongolia Faveoloolithidae Faveoloolithus Clutch(es) Coarse NA Bayinwulashan Zhao and Ding (1976); ningxiaensis cahanaobao, Alxa Left Zhang (2010) Banner (Inner Mongolia), Ningxia, China

219

Ootaxon/taxon Material Lithology Formation Locality References F. ningxiaensis Clutch(es) Coarse Baruungoyot Ologoy-Ulan-Tsav, Sochava (1969); Gobi, Mongolia Mikhailov (1994a) F. sp. Clutch(es) Fine* Seonso Bosung County, Huh and Zelenitsky Chullanam-do, South (2002); Kim et al. (2009); Korea Paik et al. (2004) Parafaveoloolithus Clutch(es) Fine Laijia Laijia Village, Tiantai, Zhang (2010) macroporus Zhejiang, China Faveoloolithidae indet. Clutch(es) Coarse Shiwa Shiwa Bay, Whaseong, Lee (2003) Gyeonggi, South Korea Faveoloolithidae indet. Clutch(es) Fine Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Locations 1-1 I5, South Korea 1-2, and 1-4) Faveoloolithidae indet. Clutch(es) Fine Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 3-1 South Korea Islets between) Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Locations 3-2 HH South Korea and JH, 3-3, and 3-4) Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 3-5) South Korea

220

Ootaxon/taxon Material Lithology Formation Locality References Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 3-6) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Locations 4-1 JH, South Korea 4-2, 4-2, 4-3, and 4-4) Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 4-5) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 5-2) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 5-3) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 5-4) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Locations 6-1 H South Korea and 6-2) Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Locations 6-3 and South Korea 6-4) Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009)

221

Ootaxon/taxon Material Lithology Formation Locality References (Location 6-5) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 6-6) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 6-7) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 6-8) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 6-9) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 9-1 D) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 9-2) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 9-3) South Korea Faveoloolithidae indet. Clutch(es) Coarse Sihwa Hanyom, Sihwa Basin, Kim et al. (2009) (Location 9-4) South Korea Faveoloolithidae indet. Clutch(es) Coarse Haman Coast of Sinsudo Island, Paik et al. (2012) Samcheonpo, South Korea Faveoloolithidae indet. Colonial nesting Coarse Los Llanos Sanagasta, La Rioja, Grellet-Tinner and

222

Ootaxon/taxon Material Lithology Formation Locality References ground (one of 5 Argentina Fiorelli (2010); Fiorelli et horisons) al. (2012) Faveoloolithidae indet. Colonial nesting Coarse* Los Llanos Sanagasta, La Rioja, Grellet-Tinner and ground (one of 5 Argentina Fiorelli (2010); Fiorelli et horisons) al. (2012) Faveoloolithidae indet. Colonial nesting Coarse* Los Llanos Sanagasta, La Rioja, Grellet-Tinner and ground (one of 5 Argentina Fiorelli (2010); Fiorelli et horisons) al. (2012) Faveoloolithidae indet. Colonial nesting Coarse* Los Llanos Sanagasta, La Rioja, Grellet-Tinner and ground (one of 5 Argentina Fiorelli (2010); Fiorelli et horisons) al. (2012) Faveoloolithidae indet. Colonial nesting Coarse* Los Llanos Sanagasta, La Rioja, Grellet-Tinner and ground (one of 5 Argentina Fiorelli (2010); Fiorelli et horisons) al. (2012) Faveoloolithidae indet. Clutch(es) (Egg Coarse* Allen Arriagada I, Rio Negro, Salgado et al. (2007) bed 3 of Salitral Argentina Ojo de Agua) Faveoloolithidae indet. Clutch(es) (Egg Coarse* Allen Santos I, Salitral de Salgado et al. (2007) bed 2 of Salitral de Santa Rosa-Salinas de Santa Rosa-Salinas Trapalco, Rio Negro, de Trapalco) Argentina

223

Ootaxon/taxon Material Lithology Formation Locality References Faveoloolithidae indet. Clutch(es) (Egg Coarse* Allen Santos II and Berthe IV, Salgado et al. (2007) bed 3 of Salitral de Salitral de Santa Santa Rosa-Salinas Rosa-Salinas de de Trapalco) Trapalco, Rio Negro, Argentina Faveoloolithidae indet. Clutch(es) Coarse NA Algui Ulan Tsav, Gobi, Watabe et al. (2010) Mongolia Megaloolithidae Megaloolithus Clutch(es) Coarse* Lameta Balasinor Quarry, Mohabey (1998) balasinorensis Jetholi and Dhuvadiya, (/baghensis) Kheda, Gujarat, India M. dhoridungriensis Clutch(es) with Coarse* Lameta Dhoridungri, Gujarat, Mohabey (1998) juvenile bones India M. dhoridungriensis Clutch(es) with Coarse* Lameta Near Dhoridungri, Wilson et al. (2010) hatchling bones Gujarat, India M. Clutch(es) Coarse* Lameta Khempur and Werasa, Mohabey (1998) khempurensis(/walpurens Kheda, Gujarat, India is?) M. mammilare Clutch(es) (one of Fine NA Commune de Rousset, Williams et al. (1984) multiple near Cairanne, stratigraphic Aix-en-Provence,

224

Ootaxon/taxon Material Lithology Formation Locality References levels) France

M. mammilare Clutch(es) (one of Fine NA Commune de Williams et al. (1984) multiple Chateauneuf Le Rouge, stratigraphic Aix-en-Provence, levels) France M. matleyi Clutch(es) (Site B) Coarse* Lameta Pavna, Chandrapur, Mohabey (1996, 1998) (/jabalpurensis) Maharashtra, India M. matleyi Nesting ground Coarse* Lameta Pavna, Chandrapur, Mohabey (1996) (/jabalpurensis) (Site A) Maharashtra, India M. megadermus Clutch(es) Coarse* Lameta Dholidhanti and Paori, Mohabey (1998) Dohad, Panchmahals, India M. microtuberculata Clutch(es) Fine NA La Cairanne, Garcia and Vianey-Liaud Aix-en-Provence, (2001); Vianey-Liaud et France al. (2003) M. patagonicus Colonial nesting Fine* Anacleto Auca Mahuevo, Chiappe et al. (1998, (Titanosauridae) ground (Egg bed 3) Neuquen, Patagonia, 2001); Grellet-Tinner et with embryonic Argentina al. (2004); Jackson et al. bones (2013) M. patagonicus Colonial nesting Fine* Anacleto Auca Mahuevo, Chiappe et al. (2001,

225

Ootaxon/taxon Material Lithology Formation Locality References (Titanosauridae) ground (Egg bed 4) Neuquen, Patagonia, 2004) Argentina M. patagonicus Colonial nesting Fine Anacleto Auca Mahuevo, Chiappe et al. (2001, (Titanosauridae) ground (Egg bed 2) Neuquen, Patagonia, 2004) Argentina M. patagonicus Colonial nesting Fine Anacleto Auca Mahuevo, Chiappe et al. (2001, (Titanosauridae) ground (Egg bed 1) Neuquen, Patagonia, 2004) Argentina M. patagonicus Clutch(es) Fine Rio Colorado Gran Neuquen Calvo et al. (1997) (Titanosauridae) neighbourhood, Neuquen, Patagonia, Argentina M. phensaniensis Clutch(es) Coarse* Lameta Phensani, Balasinor, Mohabey (1998) (/mohabeyi) Sonipur and Waniawao, Kheda, Gujarat, India M. rahioliensis Clutch(es) Coarse* Lameta Rahioli, Kheda, Gujarat, Mohabey (1998) (/cylindricus) India M. siruguei Clutch(es) (Egg Fine* Tremp Coll de Nargo, Lleida, Jackson (2007); Vila et horizon 1) Catalonia, Spain al. (2010) M. siruguei Clutch(es) (Egg Fine* Tremp Coll de Nargo, Lleida, Jackson (2007); Vila et horizon 3) Catalonia, Spain al. (2010)

226

Ootaxon/taxon Material Lithology Formation Locality References M. siruguei Clutch(es) (Egg Fine* Tremp Coll de Nargo, Lleida, Jackson (2007); Vila et horizon 5) Catalonia, Spain al. (2010) M. siruguei Clutch(es) (Egg Coarse* Tremp Coll de Nargo, Lleida, Jackson (2007); Vila et horizon 6) Catalonia, Spain al. (2010) M. siruguei Clutch(es) (Egg Coarse* Tremp Coll de Nargo, Lleida, Jackson (2007); Vila et horizon 7) Catalonia, Spain al. (2010) M. siruguei Clutch(es) (Egg Coarse* Tremp Coll de Nargo, Lleida, Jackson (2007); Vila et horizon 8) Catalonia, Spain al. (2010) M. siruguei Clutch(es) Fine* Tremp Font del Bullidor, Vila et al. (2010) Vallcebre Syncline, Spain M. siruguei Clutch(es) Fine NA Commune de Williams et al. (1984) -Amphoux, near Metisson, Aix-en-Provence, France M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013)

227

Ootaxon/taxon Material Lithology Formation Locality References multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Coarse Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain

228

Ootaxon/taxon Material Lithology Formation Locality References stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Fine* Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Clutch(es) (one of Coarse Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic

229

Ootaxon/taxon Material Lithology Formation Locality References levels) M. siruguei Clutch(es) Fine Tremp Biscarri, Isona, Lleida, Lopez-Martinez et al. Spain (2000) M. cf. siruguei Clutch(es) Fine Densus-Ciula Tustea, Hateg Basin, Grigorescu et al. (1994, Romania 2010) M. cf. siruguei Clutch(es) Fine* Sanpetru Totesti-baraj, Hateg Codrea et al. (2002); (Horizon E) Basin, Romania Grigorescu et al. (1994, 2010); Grellet-Tinner et al. (2012) M. cf. siruguei Clutch(es) Fine* Sanpetru Totesti-baraj, Hateg Codrea et al. (2002); (Horizon A) Basin, Romania Grigorescu et al. (1994, 2010); Grellet-Tinner et al. (2012) M. cf. siruguei Clutch(es) Fine* Sanpetru Totesti-baraj, Hateg Codrea et al. (2002); (Horizon K) Basin, Romania Grigorescu et al. (1994, 2010); Grellet-Tinner et al. (2012) M. cf. siruguei Clutch(es) Fine* Sanpetru Totesti-baraj, Hateg Codrea et al. (2002); (Horizon B) Basin, Romania Grigorescu et al. (1994, 2010); Grellet-Tinner et al. (2012)

230

Ootaxon/taxon Material Lithology Formation Locality References M. cf. siruguei Clutch(es) Fine* Sanpetru Totesti-baraj, Hateg Codrea et al. (2002); (Horizons G and Basin, Romania Grigorescu et al. (1994, H) 2010); Grellet-Tinner et al. (2012) M. cf. siruguei Clutch(es) (Find Fine* Sanpetru Nalat-Vad, Hateg Basin, Smith et al. (2002); 8) Romania Grigorescu et al. (1994, 2010) M. cf. siruguei Clutch(es) (Find Fine* Sanpetru Nalat-Vad, Hateg Basin, Smith et al. (2002); 9) Romania Grigorescu et al. (1994, 2010) M. cf. siruguei Clutch(es) (Find Fine* Sanpetru Nalat-Vad, Hateg Basin, Smith et al. (2002); 2) Romania Grigorescu et al. (1994, 2010) M. cf. siruguei Clutch(es) (Find Fine Sanpetru Nalat-Vad, Hateg Basin, Smith et al. (2002); 6) Romania Grigorescu et al. (1994, 2010) M. cf. siruguei Clutch(es) (Find Coarse* Sanpetru Nalat-Vad, Hateg Basin, Smith et al. (2002); 7) Romania Grigorescu et al. (1994, 2010) M. sp. Clutch(es) Coarse* Lameta Salbardi, District Srivastava and Mankar Amravati, Maharashtra (2015)

231

Ootaxon/taxon Material Lithology Formation Locality References and Betul, India M. sp. Colonial nesting Coarse Aren Barranc de la Costa Sanz et al. (1995); Sander ground Sandstone Gran, Basturs, Tremp et al. (1998); syncline, southern Diaz-Molina et al. (2007) Pyrenees, Spain M. sp. Clutch(es) Fine Tremp Coll de Nargo, Lleida, Sander et al. (1998); (Horizon 1) Catalonia, Spain Peitz (2001) M. sp. Clutch(es) Fine Tremp Coll de Nargo, Lleida, Sander et al. (1998); (Horizon 2) Catalonia, Spain Peitz (2001) M. sp. Clutch(es) Fine Tremp Coll de Nargo, Lleida, Sander et al. (1998); (Horizon 3) Catalonia, Spain Peitz (2001) M. sp. Clutch(es) Fine Tremp Coll de Nargo, Lleida, Sander et al. (1998); (Horizon 4) Catalonia, Spain Peitz (2001) M. sp. Clutch(es) Fine Tremp Coll de Nargo, Lleida, Sander et al. (1998); (Horizon 5) Catalonia, Spain Peitz (2001) M. sp. Clutch(es) Fine Tremp Coll de Nargo, Lleida, Sander et al. (1998); (Horizon 6) Catalonia, Spain Peitz (2001) M. sp. Clutch(es) Fine Tremp Coll de Nargo, Lleida, Sander et al. (1998); (Horizon 7) Catalonia, Spain Peitz (2001) M. sp. Clutch(es) Fine* Densus-Ciula Tustea, Hateg Basin, Venczel et al. (2015) Romania

232

Ootaxon/taxon Material Lithology Formation Locality References M. sp. (Type 2A: Clutch(es) (Egg Coarse* Allen Berthe IV, Salitral de Salgado et al. (2007, probably M. patagonicus) level 3 of Salitral Santa Rosa-Salinas de 2009) de Santa Trapalco, Rio Negro, Rosa-Salinas de Argentina Trapalco) Megaloolithidae indet. Clutch(es) Coarse* Allen Bajo de Arriagada, Genise and Sarzetti Patagonia, Rio Negro, (2011) Argentina Megaloolithidae indet. Clutch(es) (Egg Coarse Allen Arrigada III, Salitral Ojo Salgado et al. (2007) (Type 2B) level 2 of Salitral de Agua, Rio Negro, Ojo de Agua) Argentina Megaloolithidae indet. Clutch(es) Coarse Allen Arrigada III, Rio Negro, Salgado et al. (2007) Argentina Megaloolithidae indet. Clutch(es) Coarse Los Llanos Tama, Sierra de Los Hechenleitner et al. Llanos, La Rioja, (2016) Argentina "Hypselosaurus Colonial nesting Fine NA (Marines Rennes-le-chateu, Aude, Cousin et al. (1989) priscus" ground Rouges France Inferieures) Montanoolithidae Montanoolithus Clutch(es) Coarse* Two Medicine Blackfeet Reservation, Zelenitsky and Therrien

233

Ootaxon/taxon Material Lithology Formation Locality References strongorum Montana, USA (2008b) Ovaloolithidae Ovaloolithus sp. Clutch(es) Fine Chichengshan Ganjiechang, Tiantai, Fang et al. (2000) Zhejiang, China Ovaloolithidae indet. Clutch(es) Fine* Hasandong Sumunri, Lee (2003); Yun and Geumseong-myeon, Yang (1997); Paik et al. Hadong, South (2012) Gyeongsang, South Korea Preprismatoolithus Preprismatoolithus Clutch(es) Fine Morrison Young Locality, Delta Hirsch (1994) coloradensis (Allosaurus) County, Colorado, USA

Preprismatoolithus Clutch(es) with Fine Lourihna Pai Mogo, Lourinha, Mateus et al. (1997); (Lourinhanosaurus embryonic bones Estramadur, Portugal Antunes et al. (1998) antunesi) Prismatoolithidae Prismatoolithus Clutch(es) Fine Gaogou Xixia Basin, Henan, Wang and Zhou (1995) gebiensis China P. gebiensis Clutch(es) Fine Djadokhta Bayan Mandahu, Urad Zhao and Li (1993) Houqi, Inner Mongolia,

234

Ootaxon/taxon Material Lithology Formation Locality References China P. heyuanensis Clutch(es) Coarse Dongyuan Fengguang Village of Lü et al. (2006); Tanaka Yuannan Town, et al. (2012) Yuancheng District, Heyuan, Guangdong, China P. jenseni Clutch(es) Coarse North Horn Sauropod Locality, Jensen (1966); Bray Emery County, Utah, (1999) USA P. levis (Troodon Clutch(es) Fine* Oldman New Egg Site, Devil's Zelenitsky and Hills formosus) Coulee, Alberta, Canada (1996); this study (TMP 1994.179.1) P. levis (Troodon Clutch(es) Fine* Oldman North Baby Butte, This study (TMP formosus) Devil's Coulee, Alberta, 1996.86.1) Canada P. levis (Troodon Clutch(es) Fine* Two Medicine Egg Mountain, Willow Varricchio et al. (1997, formosus) Creek Anticline and the 1999, 2002) Red Rock Locality, Choteau, Teton County, Montana, USA P. levis (Troodon Clutch(es) Fine Two Medicine Unnamed site, Willow Varricchio et al. (2002,

235

Ootaxon/taxon Material Lithology Formation Locality References formosus) Creek Anticline and the 2015) Red Rock Locality, Choteau, Teton County, Montana; Glacier County, Montana, USA P. tiantaiensis Clutch(es) Fine Chichengshan Tiantai Basin, Zhejiang, Qian et al. (2008) China Protoceratopsidovum Clutch(es) Coarse Barun-Goyot Khermin-Tsav, Gobi, Mikhailov (1994b) fluxuosum Mongolia P. minimum Clutch(es) Coarse Djadokhta Baga-Tariach, Gobi, Mikhailov (1994b) Mongolia P. sincerum Clutch(es) Coarse Djadokhta Bain-Dzak, Gobi, Mikhailov (1994b) Mongolia P. sincerum Clutch(es) Coarse Djadokhta Main Field, Bain-Dzak, Sabath (1991) Gobi, Mongolia P. sincerum Clutch(es) Coarse Djadokhta Dashzeveg Sayr, Sabath (1991) Bain-Dzak, Gobi, Mongolia P. sp. Clutch(es) Coarse* Djadokhta Zhinst Tolgoi, Udyn Suzuki and Watabe Sayr, Gobi, Mongolia (2000a) P. sp. Clutch(es) Coarse Djadokhta Bain-Dzak, Gobi, Suzuki and Watabe

236

Ootaxon/taxon Material Lithology Formation Locality References Mongolia (2000b) P. sp. Clutch(es) Coarse Djadokhta Tugrikin Shire, Gobi, Watabe et al. (2010) Mongolia Sankofa pyrenaica Clutch(es) Coarse Aren Serrat Pedrego, Montsec Lopez-Martinez and area, Lleida, Catalonia, Vicens (2012) Spain Prismatoolithus sp. Clutch(es) Fine* Oldman Devil's Coulee, Alberta, This study (TMP Canada 2008.75.51) P. levis (Troodon Clutch(es) with Fine* Two Medicine Egg Island, Choteau, Varricchio et al. (2002) formosus) embryonic bones Teton County, Montana, USA Spheroolithidae Paraspheroolithus Clutch(es) Coarse Dongyuan Heyuan Basin, Fang et al. (2005) sanwangbacunensis Guangdong, China Clutch(es) Fine Wangshi Near Chinkongkou and Chow (1951, 1954) spheroides Group Tsotan, Laiyang, Shangdong, China S. sp. Clutch(es) (one Fine* Seonso Bosung County, Huh and Zelenitsky from at least two Chullanam-do, South (2002); Kim et al. (2009); stratigraphic Korea Paik et al. (2004) levels)

237

Ootaxon/taxon Material Lithology Formation Locality References S. sp. Clutch(es) (one Fine* Seonso Bosung County, Huh and Zelenitsky from at least two Chullanam-do, South (2002); Kim et al. (2009); stratigraphic Korea Paik et al. (2004) levels) Spheroolithidae indet. Clutch(es) (Site 2- Fine Goseong Goseong, South Korea Paik et al. (2012) lower horizon) Spheroolithidae indet. Clutch(es) (Site 3) Fine* Goseong Goseong, South Korea Paik et al. (2012) Spheroolithidae indet. Clutch(es) (Site 4) Fine Goseong Goseong, South Korea Paik et al. (2012) Spheroolithidae indet. Clutch(es) Coarse Barungoyot Shiluut Ula, Gobi, Watabe et al. (2010) Mongolia Spheroolithidae indet. Clutch(es) Coarse* Djadokhta Dzamin Khond, Gobi, Watabe et al. (2010) Mongolia Maiasaura peeblesorum Colonial nesting Fine* Two Medicine Choteau, Teton County, Horner and Makela ground Montana, USA (1979); Horner (1982) Hypacrosaurus Clutch(es) Fine* Two Medicine Blacktail Creek, Glacier Horner and Currie (1994) stebingeri County, Montana, USA H. stebingeri Clutch(es) (one of Fine* Oldman Diablo's Hill, Devil's Horner and Currie multiple Coulee, Alberta, Canada (1994); this study (TMP stratigraphic 1989.79.53) levels) H. stebingeri Clutch(es) with Fine* Oldman Diablo's Hill, Devil's Horner and Currie

238

Ootaxon/taxon Material Lithology Formation Locality References embryonic bones Coulee, Alberta, Canada (1994); this study (TMP (one of multiple 1988.79.36) stratigraphic levels) H. stebingeri Clutch(es) Fine* Oldman Diablo's Hill, Devil's This study (TMP Coulee, Alberta, Canada 1997.63.1) Youngoolithidae Youngoolithus Clutch(es) Fine NA Xianguan Basin, Zhao (1979a); Zhang xiaguanensis Neixiang County, (2010) Henan, China Oofamily incertae sedis Continuoolithus Clutch(es) Fine* Two Medicine Flaming Cliff locality, Jackson et al. (2015) canadensis Willow Creek Anticline and the Red Rock Locality, Choteau, Teton County, Montana, USA Clutch(es) (one of Fine* Upper Elliot Golden Gate Highlands Kitching (1979); carinatus four horisons) National Park, South Zelenitsky and Modesto Africa (2002); Reisz et al. (2012)

239

Ootaxon/taxon Material Lithology Formation Locality References M. carinatus Clutch(es) (one of Fine* Upper Elliot Golden Gate Highlands Kitching (1979); four horisons) National Park, South Zelenitsky and Modesto Africa (2002); Reisz et al. (2012) M. carinatus Clutch(es) (one of Fine* Upper Elliot Golden Gate Highlands Kitching (1979); four horisons) National Park, South Zelenitsky and Modesto Africa (2002); Reisz et al. (2012) M. carinatus Clutch(es) (one of Fine* Upper Elliot Golden Gate Highlands Kitching (1979); four horisons) National Park, South Zelenitsky and Modesto Africa (2002); Reisz et al. (2012) Multicanaliculate eggs Clutch(es) Coarse* Mercedes Soriano, Uruguay Faccio (1994) Indeterminate dinosaur Clutch(es) Coarse Jindong Goseong, South Korea Lee (2003) eggs

240

Appendix 4.4. Lithology of dinosaur eggs and eggshells that are unlikely in-situ (assumedly ex-situ). Note that * indicates that additional lithologic data were retrieved from the collection database at the Royal Tyrrell Museum of Palaeontology, Drumheller, Canada. Abbreviation: NA, not available. Ootaxon/taxon Material Lithology Formation Locality References Arriagadoolithidae Arriagadoolithus Isolated egg(s) with Coarse Allen Salitral Ojo de Agua, Agnolin et al. (2012) patagoniensis Bonapartenykus North Patagonia, (Bonapartenykus ultimus?) bones Argentina Cairanoolithidae Cairanoolithus dughii Eggshell(s) Fine Red Marls Les Boudous and Fondevilla et al. (2016) of Maurine Gourg de l'Encantado, Formation Upper Aude Valley, France C. cf. roussetensis Isolated egg(s) (one Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) of multiple Catalonia, Spain stratigraphic levels) C. cf. roussetensis Isolated egg(s) (one Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) of multiple Catalonia, Spain stratigraphic levels) Dendroolithidae (/Phaceloolithidae) Dendroolithus dendriticus Eggshell(s) Fine Zhaoying Xinping-Chimei Fang et al. (2007)

241

Ootaxon/taxon Material Lithology Formation Locality References Basin, Xixia, Henan, China Phaceloolithidae indet. Isolated egg(s) Coarse Lourinha Porto das Barcas, near Ribeiro et al. (2014) Lourinha, Portugal Dendroolithidae indet. Eggshell(s) Coarse Barungoyot/ Khermeen Tsav, Watabe et al. (2010) Nemegt Gobi, Mongolia Dendroolithidae indet. Eggshell(s) Coarse ?Nemegt Yagaan Khovil, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae Elongatoolithus andrewsi Eggshell(s) (one of Fine Zhutian Nanxiong Basin, Fang et al. (2009) multiple Guangdong, China stratigraphic levels) E. andrewsi Eggshell(s) (one of Fine Zhutian Nanxiong Basin, Fang et al. (2009) multiple Guangdong, China stratigraphic levels) E. andrewsi Eggshell(s) (one of Fine Zhutian Nanxiong Basin, Fang et al. (2009) multiple Guangdong, China stratigraphic levels) E. sp. Eggshell(s) Fine Fenshuiao Western Dongting Zeng and Zhang (1979) Basin, Hunan, China E. sp. Eggshell(s) Coarse ?Djadokhta Toosgot area, Abdrant Watabe (2004)

242

Ootaxon/taxon Material Lithology Formation Locality References Nuru, Gobi, Mongolia E. sp. Eggshell(s) Fine 'Lower Kamitaki, Tamba, Tanaka et al. (2016) Formation' Hyogo, Japan of Sasayama Group E. sp. Eggshell(s) Coarse Nemegt Khaichin Ula, Gobi, Suzuki and Watabe Mongolia (2000b) Macroelongatoolithus sp. Eggshell fragments Coarse Wayan Robinson Bonebed Krumenacker et al. (IMNH2251), (2016) Bonneville County, Idaho, USA Macroelongatoolithus sp. Eggshell fragments Fine Wayan Brockman Creek Krumenacker et al. (IMNH2427), (2016) Bonneville County, Idaho, USA Macroelongatoolithus sp. Eggshell fragments Fine Wayan Jackknife Creek Krumenacker et al. (IMNH2428), (2016) Bonneville County, Idaho, USA Macroolithus yaotunensis Isolated egg(s) Fine Pingling Datang, Nanxiong Zhao et al. (2002) Basin, Guangdong,

243

Ootaxon/taxon Material Lithology Formation Locality References China Megafusoolithus Isolated egg(s) Fine Chichengsh Shuangtang, Tiantai Wang et al. (2010) qiaoxiaensis an Basin, Zhejiang, China Paraelongatoolithus Isolated egg(s) Fine Chichengsh Chengguan, Tiantai Wang et al. (2010) reticulatus an Basin, Zhejiang, China Trachoolithus faticanus Eggshell(s) Fine Dushi Ula Buylyasutuin-Khuduk Kurzanov and Mikhailov (Doshuul) , Ubur-Khangay, (1989); Mikhailov Mongolia (1994b) Elongatoolithidae indet. Eggshell(s) (Egg Coarse Allen Arrigada III, Salitral Salgado et al. (2007) level 2 of Salitral Ojo de Agua, Rio Ojo de Agua) Negro, Argentina Elongatoolithidae indet. Eggshell(s) Fine Quipa Luanchuan, Tantou Xu (2007); Tanaka et al. Basin, Henan, China (2011); Jiang et al. (2011) Elongatoolithidae indet. Isolated egg(s) Fine Dadaepo Busan, South Korea Paik et al. (2012) Elongatoolithidae indet. Eggshell(s) Fine Blesa La Cantalera, Teruel, Canudo et al. (2010) Spain Elongatoolithidae indet. Eggshell(s) Coarse Djadokhta Dzamin Khond, Gobi, Suzuki and Watabe Mongolia (2000b)

244

Ootaxon/taxon Material Lithology Formation Locality References Elongatoolithidae indet. Eggshell(s) Coarse Djadokhta Bayn Dzak, Gobi, Suzuki and Watabe Mongolia (2000b) Elongatoolithidae indet. Eggshell(s) Coarse Djadokhta Mongot, Gobi, Suzuki and Watabe Mongolia (2000b) Elongatoolithidae indet. Isolated egg(s) Coarse Nemegt Bugin Tsav, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae indet. Isolated egg(s) (Red Coarse Barungoyot/ Khermeen Tsav, Watabe et al. (2010) Beds) Nemegt Gobi, Mongolia Elongatoolithidae indet. Eggshell(s) (White Coarse Barungoyot/ Khermeen Tsav, Watabe et al. (2010) Beds) Nemegt Gobi, Mongolia Elongatoolithidae indet. Eggshell(s) Coarse Nemegt Khermeen Tsav-II, Watabe et al. (2010) Gobi, Mongolia Elongatoolithidae indet. Eggshell(s) Fine Nemegt Shar Tsav, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae indet. Eggshell(s) Coarse Djadokhta Tugrikin Shire, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae indet. Eggshell(s) Coarse Djadokhta Dzamin Khond, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae indet. Eggshell(s) Coarse Djadokhta Bortolgoi, Gobi, Watabe et al. (2010) Mongolia Elongatoolithidae indet. Eggshell(s) Coarse ?Nemegt Yagaan Khovil, Gobi, Watabe et al. (2010)

245

Ootaxon/taxon Material Lithology Formation Locality References Mongolia Faveoloolithidae Faveoloolithus sp. Isolated egg(s) Fine Seonso Bosung County, Huh and Zelenitsky Chullanam-do, South (2002); Kim et al. Korea (2009); Paik et al. (2004) F. sp. Isolated egg(s) Fine Seonso Bosung County, Huh and Zelenitsky Chullanam-do, South (2002); Kim et al. Korea (2009); Paik et al. (2004) Parafaveoloolithus Isolated egg(s) Fine Laijia Fangshan, Tiantai, Zhang (2010) microporus Zhejiang, China Sphaerovum erbei Eggshell(s) Coarse Asencio Colonia, 12a. Police Mones (1980) Section, stream Miguelete Tala, Uruguay Faveoloolithidae indet. Isolated egg(s) Fine Sihwa Hanyom, Sihwa Kim et al. (2009) (Location 1-3) Basin, South Korea Faveoloolithidae indet. Isolated egg(s) Fine Sihwa Hanyom, Sihwa Kim et al. (2009) (Location 2 HH) Basin, South Korea Faveoloolithidae indet. Isolated egg(s) Fine Sihwa Hanyom, Sihwa Kim et al. (2009) (Location 5-1 SH) Basin, South Korea Faveoloolithidae indet. Isolated egg(s) Coarse Sihwa Hanyom, Sihwa Kim et al. (2009)

246

Ootaxon/taxon Material Lithology Formation Locality References (Location 6-2) Basin, South Korea Faveoloolithidae indet. Isolated egg(s) Fine Sihwa Hanyom, Sihwa Kim et al. (2009) (Location 6-10) Basin, South Korea Faveoloolithidae indet. Eggshell(s) (Egg Coarse Allen Arriagada II, Salitral Salgado et al. (2007) level 2 of Salitral Ojo de Agua, Rio Ojo de Agua) Negro, Argentina Faveoloolithidae indet. Eggshell(s) (Egg Coarse Allen Cerro de Guerra, Salgado et al. (2007) level 5 of Salitral Salitral Ojo de Agua, Ojo de Agua) Rio Negro, Argentina Faveoloolithidae indet. Eggshell(s) (Egg Coarse Allen Berthe II, Salitral de Salgado et al. (2007) level 1 of Salitral de Santa Rosa-Salinas de Santa Rosa-Salinas Trapalco, Rio Negro, de Trapalco) Argentina Faveoloolithidae indet. Eggshell(s) (Egg Coarse Allen Santos IV, Cerro Salgado et al. (2007) level 2 of Salitral de Bonaparte, Berthe I, Santa Rosa-Salinas and Berthe III, Salitral de Trapalco) de Santa Rosa-Salinas de Trapalco, Rio Negro, Argentina Faveoloolithidae indet. Eggshell(s) (Egg Coarse Allen Cerro Bonaparte, Salgado et al. (2007) level 3 of Salitral de Salitral de Santa

247

Ootaxon/taxon Material Lithology Formation Locality References Santa Rosa-Salinas Rosa-Salinas de de Trapalco) Trapalco, Rio Negro, Argentina Faveoloolithidae indet. Eggshell(s) (Egg Coarse Allen Berthe V, Berthe VI, Salgado et al. (2007) level 4 of Salitral de and Cerro Tortugas, Santa Rosa-Salinas Salitral de Santa de Trapalco) Rosa-Salinas de Trapalco, Rio Negro, Argentina Faveoloolithidae indet. Eggshell(s) (Egg Coarse Allen Garcia I, Garcia II, Salgado et al. (2007) level 5 of Salitral de and Cerro Tortugas, Santa Rosa-Salinas Salitral de Santa de Trapalco) Rosa-Salinas de Trapalco, Rio Negro, Argentina Megaloolithidae Megaloolithus aureliensis Isolated egg(s) Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) Catalonia, Spain cf. M. aureliensis Eggshell(s) Coarse Tremp Fontllonga 6, Tremp Vianey-Liaud and Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida,

248

Ootaxon/taxon Material Lithology Formation Locality References Spain M. baghensis Eggshell(s) Fine Lameta Anjar, Gujarat, India Khosla and Sahni (1995) M. cf. baghensis Isolated egg(s) Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) Catalonia, Spain M. magharebiensis Eggshell(s) Fine Irbzer Oukdiksou syncline, Garcia et al. (2003) Morocco M. mammilare Eggshell(s) Fine Red Marls Roque Fumade, Upper Fondevilla et al. (2016) of Maurine Aude Valley, France Formation Possibly M. mammillare Isolated egg(s) Coarse Tremp Bastus, Tremp Basin, Vianey-Liaud and Southern Pyrenees, Lopez-Martinez (1997) Lleida, Spain M. mammillare? Isolated egg(s) Coarse Tremp Abella, Tremp Basin, Vianey-Liaud and Southern Pyrenees, Lopez-Martinez (1997) Lleida, Spain M. matleyi (/jabalpurensis) Eggshell(s) Coarse Lameta Patbaba ridge, Mohabey (1998) Jabalpur, Madhya Pradesh M. cf. patagonicus Eggshell(s) Fine Tremp Costa de la Coma, Bravo and Gaete (2015) (Titanosauridae) Figuerola d'Orcau, Isona I Conca Della,

249

Ootaxon/taxon Material Lithology Formation Locality References South-Central Pyrenees, Spain M. petralta Eggshell(s) Coarse Tremp Fontllonga 6, Tremp Vianey-Liaud and Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida, Spain M. petralta Eggshell(s) Coarse Aren Moro, Tremp Basin, Vianey-Liaud and Southern Pyrenees, Lopez-Martinez (1997) Lleida, Spain M. pseudomamillare? (or Isolated egg(s) Coarse Tremp Suterranya, Tremp Vianey-Liaud and M. petralta) Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida, Spain M. siruguei Eggshell(s) Fine Aix-en-Prov Rousset-Routiers, Vianey-Liaud et al. ence Aix-en-Provence, (1994) France M. siruguei Eggshell(s) Coarse Aix-en-Prov Rousset-Village, Vianey-Liaud et al. ence Aix-en-Provence, (1994) France M. siruguei Eggshell(s) Fine Tremp Biscarri, Tremp Vianey-Liaud and Basin, Southern Lopez-Martinez (1997)

250

Ootaxon/taxon Material Lithology Formation Locality References Pyrenees, Lleida, Spain M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Coarse Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013)

251

Ootaxon/taxon Material Lithology Formation Locality References multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Coarse Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Coarse Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. siruguei Eggshell(s) (one of Fine Tremp Coll de Nargo, Lleida, Sellés et al. (2013) multiple Catalonia, Spain stratigraphic levels) M. cf. siruguei Eggshell(s) (one of Fine Densus-Ciul Livezi, Hateg Basin, Grigorescu et al. (1994, at least two a Romania 2010) eggshell-bearing levels)

252

Ootaxon/taxon Material Lithology Formation Locality References M. cf. siruguei Eggshell(s) (one of Fine Densus-Ciul Livezi, Hateg Basin, Grigorescu et al. (1994, at least two a Romania 2010) eggshell-bearing levels) M. sp. Eggshell(s) Fine NA Cruzy (Quarantes Garcia (2000) basin, Herault) and Vitrolles-Couperigne (Aix basin, Bouches-du-Rhone), France M. sp. Eggshell(s) Coarse Aren Basturs, Catalunya, Peitz (2001) Spain M. sp. Isolated egg(s) Fine Tremp Gosol, Catalunya, Peitz (2001) Spain Pseudomegaloolithus atlasi Eggshell(s) Fine Tendrara red Douar Lgara, Chassagne-Manoukian et beds Tendrara High al. (2013) Plateaus, Morocco Megaloolithidae indet. Isolated egg(s) Fine NA Commune de Williams et al. (1984) Pourcieux, Aix-en-Provence, France

253

Ootaxon/taxon Material Lithology Formation Locality References Megaloolithidae indet. Eggshell(s) (One of Coarse Galula TZ-07, 20 km west of Gottfried et al. (2004) five horizons) Formation the town of Mbeya in of Red the Mbeya District, Sandstone southwestern Group Tanzania Megaloolithidae indet. Eggshell(s) (One of Coarse Galula TZ-07, 20 km west of Gottfried et al. (2004) five horizons) Formation the town of Mbeya in of Red the Mbeya District, Sandstone southwestern Group Tanzania Megaloolithidae indet. Eggshell(s) (One of Coarse Galula TZ-07, 20 km west of Gottfried et al. (2004) five horizons) Formation the town of Mbeya in of Red the Mbeya District, Sandstone southwestern Group Tanzania Megaloolithidae indet. Eggshell(s) (One of Coarse Galula TZ-07, 20 km west of Gottfried et al. (2004) five horizons) Formation the town of Mbeya in of Red the Mbeya District, Sandstone southwestern Group Tanzania Megaloolithidae indet. Eggshell(s) (One of Coarse Galula TZ-07, 20 km west of Gottfried et al. (2004)

254

Ootaxon/taxon Material Lithology Formation Locality References five horizons) Formation the town of Mbeya in of Red the Mbeya District, Sandstone southwestern Group Tanzania Megaloolithidae indet. Eggshell(s) Coarse Marilia Peiropolis, Baruru Ribeiro (2002) Basin, Marilia County, San Paulo State, Brazil Megaloolithidae indet. Eggshell(s) Coarse Marilia Ponte Alta, Baruru Ribeiro (2002) Basin, Marilia County, San Paulo State, Brazil Megaloolithidae indet. Type 2A eggshell(s) Coarse Allen Arrigada I, Salitral Salgado et al. (2007) (Type 2A) (Egg level 4 of Ojo de Agua, Rio Salitral Ojo de Negro, Argentina Agua) Megaloolithidae indet. Type 2A eggshell(s) Coarse Allen Berthe II, Salitral de Salgado et al. (2007) (Type 2A) (Egg level 1 of Santa Rosa-Salinas de Salitral de Santa Trapalco, Rio Negro, Rosa-Salinas de Argentina Trapalco)

255

Ootaxon/taxon Material Lithology Formation Locality References Megaloolithidae indet. Type 2A eggshell(s) Coarse Allen Cerro Bonaparte, Salgado et al. (2007) (Type 2A) (Egg level 2 of Salitral de Santa Salitral de Santa Rosa-Salinas de Rosa-Salinas de Trapalco, Rio Negro, Trapalco) Argentina Megaloolithidae indet. Type 2A isolated Coarse Allen Berthe IV, Salitral de Salgado et al. (2007) (Type 2A) egg(s) (Egg level 4 Santa Rosa-Salinas de of Salitral de Santa Trapalco, Rio Negro, Rosa-Salinas de Argentina Trapalco) Megaloolithidae indet. Type 2A eggshell(s) Coarse Allen Cerro Turtugas, Salgado et al. (2007) (Type 2A) (Egg level 5 of Salitral de Santa Salitral de Santa Rosa-Salinas de Rosa-Salinas de Trapalco, Rio Negro, Trapalco) Argentina Megaloolithidae indet. Type 2B eggshell(s) Coarse Allen Berthe III, Salitral de Salgado et al. (2007) (Type 2B) (Egg level 2 of Santa Rosa-Salinas de Salitral de Santa Trapalco, Rio Negro, Rosa-Salinas de Argentina Trapalco) Megaloolithidae indet. Type 2B eggshell(s) Coarse Allen Mansilla I and II, Salgado et al. (2007)

256

Ootaxon/taxon Material Lithology Formation Locality References (Type 2B) (Egg level 3 of Salitral de Santa Salitral de Santa Rosa-Salinas de Rosa-Salinas de Trapalco, Rio Negro, Trapalco) Argentina Megaloolithidae indet. Type 2B eggshell(s) Coarse Allen Berthe V and VI, Salgado et al. (2007) (Type 2B) (Egg level 4 of Salitral de Santa Salitral de Santa Rosa-Salinas de Rosa-Salinas de Trapalco, Rio Negro, Trapalco) Argentina Megaloolithidae indet. Type 2B eggshell(s) Coarse Allen Garcia I, Salitral de Salgado et al. (2007) (Type 2B) (Egg level 5 of Santa Rosa-Salinas de Salitral de Santa Trapalco, Rio Negro, Rosa-Salinas de Argentina Trapalco) "Hypselosaurus priscus" Eggshell(s) Coarse NA Rennes-le-chateu, Cousin et al. (1989) Aude, France Ovaloolithidae Ovaloolithidae indet. Eggshell(s) Fine NA Tel Ulan Chaltsai Watabe et al. (2010) (Mogoin Daatsyn Khuduk), Mongolia Pinnatoolithae

257

Ootaxon/taxon Material Lithology Formation Locality References Lanceoloolithus Eggshell(s) (one of Coarse Zhutian Nanxiong Basin, Fang et al. (2009) xiapingensis multiple Guangdong, China stratigraphic levels) L. xiapingensis Eggshell(s) (one of Coarse Zhutian Nanxiong Basin, Fang et al. (2009) multiple Guangdong, China stratigraphic levels) Polyclonoolithidae Polyclonoolithus Isolated egg(s) Fine Hekou Yangjiagou, Xie et al. (2016) yangjiagouensis Group Lanzhou-Minhe Basin, Gansu, China Preprismatoolithus Preprismatoolithus Eggshell(s) Fine Lourinha Casal da Rola, near Ribeiro et al. (2014) (Lourinhanosaurus antunesi) Lourinha, Portugal P. coloradensis Isolated egg(s) Fine Morrison Cleveland-Lloyd Hirsch et al. (1989); (Allosaurus) Dinosaur Quarry, Hirsch (1994) Emery County, Utah, USA P. coloradensis Eggshell(s) Fine Morrison Garden Park Area, Hirsch (1994) (Allosaurus) near Canyon City, Colorado, USA Prismatoolithidae

258

Ootaxon/taxon Material Lithology Formation Locality References Prismatoolithus carboti Isolated egg(s) Fine Villeveyrac- Grande Marquise, Garcia et al. (2000) Meze Basin Herault, France P. carboti Eggshell(s) Fine Villeveyrac- Neuve, Garcia et al. (2000) Meze Basin Bouches-du-Rhone, France aff. P. matellensis Eggshell(s) Coarse Tremp Fontllonga 6, Tremp Vianey-Liaud and Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida, Spain P. tenuis Eggshell(s) Coarse Tremp Fontllonga 6, Tremp Vianey-Liaud and Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida, Spain Protoceratopsidovum Eggshell(s) Coarse NA Nemegt, Gobi, Sabath (1991) fluxuosum Mongolia P. minimum Isolated egg(s) Coarse Nemegt Altan Ula IV, Gobi, Sabath (1991) Mongolia P. sp. Eggshell(s) Coarse Djadokhta Bayn Dzak, Gobi, Suzuki and Watabe Mongolia (2000b) P. sp. Eggshell(s) Coarse Djadokhta Bayn Dzak, Gobi, Watabe et al. (2010) Mongolia

259

Ootaxon/taxon Material Lithology Formation Locality References Sankofa pyrenaica Eggshell(s) Coarse Aren Urbanizacion Lopez-Martinez and Montsec, Montsec Vicens (2012) area, Lleida, Catalonia, Spain Trigonoolithus amoae Eggshell(s) Fine Belsa La Cantalera site, Canudo et al. (2010); Josa, Teruel, northern Moreno-Azanza et al. Spain (2014b) Prismatoolithus sp. Eggshell(s) Coarse Dinosaur Bonebed BB098, Zelenitsky and Sloboda Park Dinosaur Provincial (2005) Park, Alberta P. sp. Eggshell(s) Fine Blesa La Cantalera, Teruel, Canudo et al. (2010) Spain P. sp. Eggshell(s) Coarse Fruitland NMMNH L-4010, Tanaka et al. (2011) Bisti/De-na-zin Wilderness area, San Juan Basin, New Mexico P. sp. Eggshell(s) Fine 'Lower Kamitaki, Tamba, Tanaka et al. (2016) Formation' Hyogo, Japan of Sasayama Group

260

Ootaxon/taxon Material Lithology Formation Locality References cf. Prismatoolithus Eggshell(s) Fine Late Cruzy (Quarantes Garcia (2000) Cretaceous basin, Herault) and Vitrolles-Couperigne (Aix basin, Bouches-du-Rhone), France Prismatoolithus? Eggshell(s) Fine Irbzer Oukdiksou syncline, Garcia et al. (2003) Morocco Prismatoolithidae indet. Eggshell(s) Fine 'Lower Kamitaki, Tamba, Tanaka et al. (2016) Formation' Hyogo, Japan of Sasayama Group Similifaveoolithidae Similifaveoloolithus Isolated egg(s) Fine Quantou Liufangzi, Wang et al. (2006, 2013) gongzhulingensis Gongzhuling, Jilin, China Spheroolithidae Guegoolithus turolensis Eggshell(s) Fine Blesa Galve, Teruel, Spain Moreno-Azanza et al. (2014a) G. turolensis Eggshell(s) Coarse Camarillas Galve, Teruel, Spain Moreno-Azanza et al. (2014a)

261

Ootaxon/taxon Material Lithology Formation Locality References Spheroolithus albertensis* Eggshell(s) Coarse Dinosaur Bonebed BB031, Zelenitsky and Sloboda Park Dinosaur Provincial (2005) Park, Alberta S. albertensis* Eggshell(s) Fine Dinosaur Bonebed BB104, Zelenitsky and Sloboda Park Dinosaur Provincial (2005) Park, Alberta S. europaeus Eggshell(s) Coarse Tremp Pont d'Orrit, Tremp, Sellés et al. (2014) Lleida, Catalonia, Spain S. megadermus Isolated egg(s) Fine Quantou Changtu, Liaoning, Liu et al. (2013) China S. spheroides (including S. Isolated egg(s) Fine Quantou Changtu, Liaoning, Liu et al. (2013) chiangchiugtingensis) China S. sp. Isolated egg(s) Fine Seonso Bosung County, Huh and Zelenitsky Chullanam-do, South (2002); Kim et al. Korea (2009); Paik et al. (2004) S. sp. Isolated egg(s) Fine Seonso Bosung County, Huh and Zelenitsky Chullanam-do, South (2002); Kim et al. (2009) Korea S. sp. Eggshell(s) Fine 'Lower Kamitaki, Tamba, Tanaka et al. (2016) Formation' Hyogo, Japan

262

Ootaxon/taxon Material Lithology Formation Locality References of Sasayama Group S. sp. Eggshell(s) with Coarse Nemegt Dragon's Tomb' Dewaele et al. (2015) perinatal dinosaur locality, Saurolophus Gobi, Mongolia angustirostris bones Spheroolithus? Eggshell(s) (one of Fine Lameta Dongargaon, Pisdura, Mohabey (1996, 2001) multiple Polgaon and Tidkepar stratigraphic levels) in Kholdoda and Chandrapur, Maharashtra, India Spheroolithus? Eggshell(s) (one of Coarse Lameta Dongargaon, Pisdura, Mohabey (1996, 2001) multiple Polgaon and Tidkepar stratigraphic levels) in Kholdoda and Chandrapur, Maharashtra, India Spheroolithidae indet. Eggshell(s) Coarse ?Djadokhta Toosgot area, Abdrant Watabe (2004) Nuru, Gobi, Mongolia Spheroolithidae indet. Eggshell(s) (Site 1: Fine Goseong Goseong, South Korea Paik et al. (2012) Lower Horizon) Spheroolithidae indet. Isolated egg(s) (Site Coarse Goseong Goseong, South Korea Paik et al. (2012)

263

Ootaxon/taxon Material Lithology Formation Locality References 1: Upper Horizon) Spheroolithidae indet. Isolated egg(s) (Site Fine Goseong Goseong, South Korea Paik et al. (2012) 2: Upper Horizon) Spheroolithidae indet. Eggshell(s) Fine Nemegt Shar Tsav, Gobi, Watabe et al. (2010) Mongolia Spheroolithidae indet. Eggshell(s) Fine Djadokhta Bortolgoi, Gobi, Watabe et al. (2010) Mongolia Spheroolithidae indet. Eggshell(s) Coarse ?Nemegt Yagaan Khovil, Gobi, Watabe et al. (2010) Mongolia cf. Spheroolithidae Eggshell(s) Fine Blesa La Cantalera, Teruel, Canudo et al. (2010) Spain Oofamily incertae sedis Ageroolithus fontllongensis Eggshell(s) Coarse Tremp Fontllonga 6, Tremp Vianey-Liaud and Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida, Spain A. fontllongensis Eggshell(s) Coarse Aren Moro, Tremp Basin, Vianey-Liaud and Southern Pyrenees, Lopez-Martinez (1997) Lleida, Spain cf. Ageroolithus Eggshell(s) Fine NA Vitrolles-Couperigne, Garcia (2000) France

264

Ootaxon/taxon Material Lithology Formation Locality References Continuoolithus canadensis Eggshell(s) Fine Oldman Little Diablo's Hill, Zelenitsky et al. (1996) Devil's Coulee, Alberta, Canada C. canadensis Eggshell(s) Fine Oldman North Baby Butte, Zelenitsky et al. (1996) Devil's Coulee, Alberta, Canada C. canadensis Eggshell(s) Fine Oldman Faye Walker's Coulee, Zelenitsky et al. (1996) Devil's Coulee, Alberta, Canada C. canadensis Eggshell(s) Fine Oldman Juvie Camp, Devil's Zelenitsky et al. (1996) Coulee, Alberta, Canada C. canadensis Eggshell(s) Fine Oldman Knight's Ranch, Zelenitsky et al. (1996) Alberta, Canada C. canadensis* Eggshell(s) Coarse Dinosaur Bonebed BB031, Zelenitsky and Sloboda Park Dinosaur Provincial (2005) Park, Alberta, Canada C. canadensis* Eggshell(s) Coarse Dinosaur Bonebed BB098, Zelenitsky and Sloboda Park Dinosaur Provincial (2005) Park, Alberta, Canada C. canadensis* Eggshell(s) Fine Dinosaur Bonebed BB104, Zelenitsky and Sloboda

265

Ootaxon/taxon Material Lithology Formation Locality References Park Dinosaur Provincial (2005) Park, Alberta, Canada C. canadensis Eggshell(s) Coarse Fruitland NMMNH L-4010, Tanaka et al. (2011) Bisti/De-na-zin Wilderness area, San Juan Basin, New Mexico, USA cf. Continuoolithus Eggshell(s) Fine Aguja "Purple Hill" Welsh and Sankey microsite, Rattlesnake (2005) Mountain, Big Bend National Park, Texas, USA Ornithoid basic type and Eggshell(s) Fine Morrison Rio Puerco drainage, Bray and Lucas (1997) ratite morphotype Arch Mesa, Sandoval (Continuoolithus?) County, New Mexico, USA antirrhopus Isolated egg(s) with Fine Cloverly 'Cashen Pocket', Crow Makovicky and embryonic Indian reservation, Grellet-Tinner (2001); Deinonychus Big Horn County, Grellet-Tinner and antirrhopus bones Montana, USA Makovicky (2006) Dispersituberoolithus exilis Eggshell(s) Fine Oldman Little Diablo's Hill, Zelenitsky et al. (1996)

266

Ootaxon/taxon Material Lithology Formation Locality References Devil's Coulee, Alberta, Canada Probably Eggshell(s) with Fine Lower Kunming, Yunnan, Reisz et al. (2013) embryonic possible Lufeng China Lufengosaurus bones Multicanaliculate eggs Eggshell(s) Coarse Mercedes Soriano, Uruguay Faccio (1994) Multicanaliculate eggs Eggshell(s) Coarse Mercedes Algorta, Uruguay Faccio (1994) Nipponoolithus ramosus Eggshell(s) Fine 'Lower Kamitaki, Tamba, Tanaka et al. (2016) Formation' Hyogo, Japan of Sasayama Group Porituberoolithus Eggshell(s) Fine Oldman Little Diablo's Hill, Zelenitsky et al. (1996) warnerensis Devil's Coulee, Alberta, Canada P. warnerensis Eggshell(s) Fine Oldman Faye Walker's Coulee, Zelenitsky et al. (1996) Devil's Coulee, Alberta, Canada P. warnerensis Eggshell(s) Fine Oldman Knight's Ranch, Zelenitsky et al. (1996) Alberta, Canada P. warnerensis* Eggshell(s) Coarse Dinosaur Bonebed BB031, Zelenitsky and Sloboda

267

Ootaxon/taxon Material Lithology Formation Locality References Park Dinosaur Provincial (2005) Park, Alberta, Canada P. warnerensis Eggshell(s) Coarse Fruitland NMMNH L-4010, Tanaka et al. (2011) Bisti/De-na-zin Wilderness area, San Juan Basin, New Mexico, USA cf. Porituberoolithus Eggshell(s) Fine Aguja "Purple Hill" Welsh and Sankey microsite, Rattlesnake (2005) Mountain, Big Bend National Park, Texas, USA Pseudogeckoolithus Eggshell(s) Coarse Tremp Fontllonga 6, Tremp Vianey-Liaud and nodosus Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida, Spain Pseudogeckoolithus? Eggshell(s) Fine NA Cruzy (Quarantes Garcia (2000) basin, Herault), France Reticuloolithus hirschi* Eggshell(s) Fine Dinosaur Bonebed BB104, Zelenitsky and Sloboda Park Dinosaur Provincial (2005)

268

Ootaxon/taxon Material Lithology Formation Locality References Park, Alberta, Canada Tristraguloolithus Eggshell(s) Fine Oldman Knight's Ranch, Zelenitsky et al. (1996) cracioides Alberta, Canada T. cracioides Eggshell(s) Fine Oldman Little Diablo's Hill, Zelenitsky et al. (1996) Devil's Coulee, Alberta, Canada T. cracioides Eggshell(s) Fine Oldman Faye Walker's Coulee, Zelenitsky et al. (1996) Devil's Coulee, Alberta, Canada Indeterminate theropod Eggshell(s) Coarse Fruitland NMMNH L-4010, Tanaka et al. (2011) eggshell Bisti/De-na-zin Wilderness area, San Juan Basin, New Mexico, USA Theropoda? Eggshell(s) Fine NA Rennes-le-chateu, Beetschen et al. (1977) Aude, France Dinosauroid spherulithic Eggshell(s) Fine Aguja "Purple Hill" Welsh and Sankey basic type microsite, Rattlesnake (2005) Mountain, Big Bend National Park, Texas, USA

269

Ootaxon/taxon Material Lithology Formation Locality References ?Prismatoolithidae Eggshell(s) Coarse Tremp Fontllonga 6, Tremp Vianey-Liaud and Basin, Southern Lopez-Martinez (1997) Pyrenees, Lleida, Spain ?Prismatoolithidae Eggshell(s) Fine Morrison Gallison Locality, Hirsch (1994) Delta County, Colorado, USA Dinosauroid prismatic basic Eggshell(s) Fine Aguja "Purple Hill" Welsh and Sankey type microsite, Rattlesnake (2005) Mountain, Big Bend National Park, Texas, USA Ratite morphotype Eggshell(s) Fine Aguja "Purple Hill" Welsh and Sankey microsite, Rattlesnake (2005) Mountain, Big Bend National Park, Texas, USA Prismatic morphotype Eggshell(s) Fine NA La Neuve and Grande Garcia (2000) Marquise, France Indet. type Isolated egg(s) Fine Oxford Clay Oxford Clay, Liston and Chapman Peterborough, (2014)

270

Ootaxon/taxon Material Lithology Formation Locality References England Indet. Type (thin ridged Eggshell(s) Fine Cloverly MOR CL-121, Maxwell and Horner type) Bridger, (1994) County, Montana, USA Indet. Type (thick ridged Eggshell(s) Fine Cloverly MOR CL-121, Maxwell and Horner type) Bridger, Carbon (1994) County, Montana, USA Indet. Type (nodes type) Eggshell(s) Fine Cloverly MOR CL-121, Maxwell and Horner Bridger, Carbon (1994) County, Montana, USA Indet. type Eggshell(s) Fine Morrison Callison Locality, Hirsch (1994) Delta County, Colorado, USA

271

Appendix 4.5. Approximate paleolatitudes of egg localities for major dinosaur ootaxa.

Abbreviations: D, Dendroolithidae; E, Elongatoolithidae; Fa, Faveoloolithidae; Me, Megaloolithidae; NA, not available; Pre,

Preprismatoolithus; Pri, Prismatoolithidae; Sp, Spheroolithidae. Positive and negative values of paleolatitudes represent north and south latitudes, respectively.

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

D Heyuan Basin, Guangdong, Dongyuan 65 29.10 Fang et al. (2005)

China

D Nanyang Valley near Xinye, Nancho 80 36.50 Manning et al. (1997);

Henan, China Kundrat et al. (2008)

D Xixia Basin (or Zhaoying 85 37.85 Fang et al. (2007)

Xinping-Chimei Basin),

Xixia, Henan, China

D Xixia Basin (or Zhoumagang 90 38.60 Fang et al. (2007)

Xinping-Chimei Basin),

Xixia, Henan, China

D Taohe, Xichuan, Henan, Majiacun 85 37.85 Zhao and Zhao (1998) 272

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

China

D Dashiqiao, Xichuan, Henan, Gaogou 90 38.40 Zhao and Zhao (1998)

China

D Yunxian, Hubei, China Gaogou 90 37.70 Zhou et al. (1998)

D Heshui, Hubei, China Gong-An-Zhai 80 35.70 Zhao and Li (1988)

D Western Dongting Basin, Fenshuiao 80 32.80 Zeng and Zhang

Hunan, China (1979)

D Tiantai Basin, Zhejiang, Chichengshan 95 36.30 Fang et al. (1998,

China 2000)

D Khermeen Tsav, Gobi, Barungoyot/ Nemegt 75 44.95 Watabe et al. (2010)

Mongolia

D Shilyust-Ula, Gobi, ?Barungoyot 75 45.20 Mikhailov (1994a)

Mongolia

D Bayn Shire, Gobi, Mongolia Baynshire 85 48.00 Watabe et al. (2010)

D Abdrant Nuru, Gobi, ?Djadokhta 70 46.60 Watabe (2004) 273

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Mongolia

D Khaichin Ula, Gobi, Nemegt 70 45.90 Suzuki and Watabe

Mongolia (2000b)

D Yagaan Khovil, Gobi, ?Nemegt 70 45.50 Watabe et al. (2010)

Mongolia

D Shiwa Bay, Whaseong, Shiwa 120 41.50 Lee (2003)

Gyeonggi, South Korea

D Porto das Barcas, near Lourinha 150 26.70 Ribeiro et al. (2014)

Lourinha, Portugal

E Maoming Basin, Tongguling 80 26.20 Zhang and Huang

Guangdong, China (1999)

E Nanxiong Basin (or Dongyuan, 65 30.35 Fang et al. (2009)

Nanxiong-Shixing Basin), Nanxiong, Pingling,

Guangdong, China Yuanpu, Zhenshui,

and Zhutian 274

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

E Sanshui-Guangzhou Basin, Sanshui 85 28.10 Zhang and Huang

Guangdong, China (1999)

E Heyuan Basin, Guangdong, Nanxiong and 65 29.10 Zhang and Huang

China Dongyuan (1999)

E Liguanqiao Basin, Henan, Sigou (or Luyemiao 75 36.80 Liang et al. (2009)

China or Hugang)

E Lingbao Basin, Henan, Sigou (or Luyemiao 75 38.35 Liang et al. (2009)

China or Hugang)

E Luanchuan, Tantou Basin, Quipa 65 38.55 Tanaka et al. (2011)

Henan, China

E Tantou Basin, Henan, China Sigou (or Luyemiao 75 38.00 Liang et al. (2009)

or Hugang)

E Xichuan Basin, Henan, Majiacun 85 37.70 Zhao and Zhao (1998)

China

E Xichuan Basin, Henan, Sigou (or Luyemiao 75 37.15 Liang et al. (2009) 275

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

China or Hugang)

E Xixia Basin (or Zhaoying 85 37.85 Fang et al. (2007)

Xinping-Chimei Basin),

Xixia, Henan, China

E Xixia Basin (or Gaogou (or 90 38.60 Liang et al. (2009)

Xinping-Chimei Basin), Zoumagang)

Xixia, Henan, China

E Western Dongting Basin, Fenshuiao 80 32.80 Zeng and Zhang

Hunan, China (1979)

E Bayan Mandahu, Inner Djadokhta 70 44.70 Dong and Currie

Mongolia, China (1996)

E Pingxiang Basin, Jiangxi, Zhoutian 80 32.00 Wang et al. (2013)

China

E Ganzhou, Jiangxi, China Nanxiong 65 31.20 Cheng et al. (2008); Ji

(2009) 276

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

E Badaohao, Heishan, Shahai 110 44.00 Zhao and Zhao (1999)

Liaoning, China

E Laiyang-Zhucheng, Jiangjunding 85 42.85 Zhao et al. (2013)

Shangdong, China Formation in the

Wangshi Group

E Laiyang-Zhucheng, Jingangkou 75 42.40 Zhao et al. (2013)

Shangdong, China Formation in the

Wangshi Group

E Tiantai Basin, Zhejiang, Chichengshan 90 35.80 Wang et al. (2010)

China

E Tiantai Basin, Zhejiang, Liangtoutang 100 36.90 Jin et al. (2007)

China

E Lavariya Muwada, Kheda, Lameta 65 -19.85 Mohabey (1998)

Gujarat, India

E Kamitaki, Tamba, Hyogo, 'Lower Formation' of 110 49.00 Tanaka et al. (2016) 277

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Japan Sasayama Group

E Northern Sayr, Nemegt, Baruungoyot 75 45.15 Fanti et al. (2012)

Gobi, Mongolia

E Khermeen Tsav, Gobi, Barungoyot/ Nemegt 75 44.95 Watabe et al. (2010)

Mongolia

E Ikh-Shunkht, Gobi, ?Barun-Goyot 75 46.60 Mikhailov (1994b)

Mongolia

E Bayn Shire, Gobi, Mongolia Baynshire 85 48.00 Watabe et al. (2010)

E Mongot, Gobi, Mongolia Djadokhta 70 46.40 Suzuki and Watabe

(2000b)

E Ukhaa Tolgod, Mongolia Djadokhta 70 45.30 Norell et al. (1994,

2001)

E Bayn Dzak, Gobi, Mongolia Djadokhta 70 46.40 Suzuki and Watabe

(2000b)

E Bortolgoi, Gobi, Mongolia Djadokhta 70 46.20 Watabe et al. (2010) 278

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

E Dzamin Khond, Gobi, Djadokhta 70 46.30 Suzuki and Watabe

Mongolia (2000b)

E Tugrikin Shire, Gobi, Djadokhta 70 46.50 Watabe et al. (2010)

Mongolia

E Abdrant Nuru, Gobi, ?Djadokhta 70 46.70 Watabe (2004)

Mongolia

E Udan-Sayr, Gobi, Mongolia Djadokhta 70 46.40 Mikhailov (1994b)

E Buylyasutuin-Khuduk, Dushi Ula (Doshuul) 120 43.10 Kurzanov and

Ubur-Khangay, Mongolia Mikhailov (1989);

Mikhailov (1994b)

E Bugin-Tsav, Gobi, Nemegt 70 45.50 Weishampel et al.

Mongolia (2008)

E Guriliyn-Tsav, Gobi, Nemegt 70 46.10 Mikhailov (1994b)

Mongolia

E Khaichin Ula, Gobi, Nemegt 70 45.90 Suzuki and Watabe 279

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Mongolia (2000b)

E Khermeen Tsav-II, Gobi, Nemegt 70 45.50 Watabe et al. (2010)

Mongolia

E Shar Tsav, Gobi, Mongolia Nemegt 70 46.90 Watabe et al. (2010)

E Tsagan-Khushu, Gobi, Nemegt 70 45.40 Mikhailov (1994b)

Mongolia

E Yagaan Khovil, Gobi, ?Nemegt 70 46.30 Watabe et al. (2010)

Mongolia

E Busan, South Korea Dadaepo 75 41.90 Paik et al. (2012)

E Tongyeong, Gyeongsang, Goseong 75 41.60 Kim et al. (2011)

South Korea

E Aphae-do, Shinan-gun, Gyeongsang Basin 80 41.00 Huh et al. (2014)

Jeollanam-do, South Korea

E Shiwa Bay, Whaseong, Shiwa 120 41.40 Lee (2003)

Gyeonggi, South Korea 280

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

E Teruel, Spain Blesa 130 25.60 Canudo et al. (2010)

E Bonneville County, Idaho, Wayan 109.0–99.7 43.30 Krumenacker et al.

USA (2016)

E Pancake and Creek Newark Canyon 125.5– 99.7 41.90 Bonde et al. (2015)

Ranges, , USA

E Castle Dale, Emery County, Cedar Mountain 105.3– 94.3 39.70 Zelenitsky et al.

Utah (2000)

E Section 4 Site 7, Kitchen North Horn 70.6–66.0 47.20 Bray (1999)

Locality, Sauropod Locality,

Sauropod One Egg, Pond

Locality, Locality,

and North Horn Mountain,

Emery County, Utah, USA

E UCM Loc. #78203, Emery Dakota 105.3–94.3 39.70 Zelenitsky et al.

County, Utah, USA (2000) 281

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

E Salitral Ojo de Agua, Rio Allen 75 -44.05 Salgado et al. (2007)

Negro, Argentina

Fa Qinglongshan, Yunxian, Gaogou 90 37.70 Zhou et al. (1998)

Hubei, China

Fa Pingxiang Basin, Jiagxi, Zhoutian 80 32.00 Zou et al. (2013)

China

Fa Tiantai, Zhejiang, China Laijia 95 36.35 Zhang (2010)

Fa Algui Ulan Tsav, Gobi, NA 115 44.25 Watabe et al. (2010)

Mongolia

Fa Ologoy-Ulan-Tsav, Gobi, Barungoyot 75 47.10 Sochava (1969);

Mongolia Mikhailov (1994a)

Fa Bosung County, Seonso 80 41.00 Huh and Zelenitsky

Chullanam-do Province, (2002)

South Korea

Fa Coast of Sinsudo Island, Haman 100 43.20 Paik et al. (2012) 282

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Samcheonpo, South Korea

Fa Hanyom, Sihwa Basin, Sihwa 120 41.50 Kim et al. (2009)

South Korea

Fa Ojo de Agua, Rio Negro, Allen 75 -44.05 Salgado et al. (2007)

Argentina

Fa Salitral de Santa Allen 75 -44.85 Salgado et al. (2007)

Rosa-Salinas de Trapalco,

Rio Negro, Argentina

Fa Sanagasta, La Rioja, Los Llanos 125 -26.10 Grellet-Tinner and

Argentina Fiorelli (2010);

Fiorelli et al. (2012)

Fa Colonia, 12a. Police Section, Asencio 95 -38.75 Mones (1980)

stream Miguelete Tala,

Uruguay

Fa Quebracho, Paysandu, Guichon 100 -36.10 Soto et al. (2012) 283

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Uruguay

Me Douar Lgara, Tendrara High Tendrara red beds 65 19.15 Chassagne-Manoukia

Plateaus, Morocco n et al. (2013)

Me Oukdiksou syncline, Irbzer 65 19.95 Garcia et al. (2003)

Morocco

Me TZ-07, Mbeya District, Galula Formation of 105 -28.70 Gottfried et al. (2004)

Tanzania Red Sandstone Group

Me Rahioli and Balasinor Lameta 65 -19.85 Srivastava et al.

region, Kheda, Gujarat, (1986)

India

Me Anjar, Kachchh, Gujarat, Lameta 65 -18.30 Khosla and Sahni

India (1995)

Me Jhabua, Madhya Pradesh, Lameta 65 -20.40 Khosla and Sahni

India (1995)

Me Khempur, Kheda, Gujarat, Lameta 65 -18.30 Khosla and Sahni 284

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

India (1995)

Me Dohad, Panchmahals, Lameta 65 -20.25 Mohabey and Mathur

Gujarat, India (1989)

Me Dhar, Madhya Pradesh, Lameta 65 -20.80 Khosla and Sahni

India (1995)

Me Padiyal, Madhya Pradesh, Lameta 65 -20.95 Khosla and Sahni

India (1995)

Me Jabalpur, Madhya Pradesh, Lameta 65 -21.55 Khosla and Sahni

India (1995)

Me Amravati, Maharashtra, Lameta 65 -23.10 Srivastava and

India Mankar (2015)

Me Chandrapur, Maharashtra, Lameta 65 -24.50 Mohabey (1996,

India 1998)

Me Durlston Bay, Swanage, Chery Freshwater 140 33.10 Ensom (1996)

Dorset, England Member of the 285

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Lulworth Formation

Me Aix-en-Provence region, NA 75 33.20 Vianey-Liaud et al.

France (1994)

Me Aix-en-Provence region, NA 65 32.45 Vianey-Liaud et al.

France (1994)

Me Aude region, France Red Marls of 65 32.00 Fondevilla et al.

Maurine (2016)

Me Hateg Basin, Romania Densus-Ciula 70.6–66.0 32.70 Grigorescu et al.

(1994, 2010)

Me Southern Pyrenees region Aren 70 30.00 Sanz et al. (1995);

(e.g., Coll de Nargo and Diaz-Molina et al.

Tremp), Spain (2007)

Me Southern Pyrenees region Tremp 65 29.65 Vianey-Liaud and

(e.g., Coll de Nargo and Lopez-Martinez

Tremp), Spain (1997) 286

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Me Auca Mahuevo, Neuquen Anacleto 80 -42.30 Chiappe et al. (1998)

Province, Patagonia,

Argentina

Me Arrigada I, Salitral Ojo de Allen 75 -44.05 Salgado et al. (2007)

Agua, Rio Negro, Argentina

Me Salitral de Santa Allen 75 -44.85 Salgado et al. (2007)

Rosa-Salinas de Trapalco,

Rio Negro, Argentina

Me Bajo de Arriagada, Allen 75 -44.00 Genise and Sarzetti

Patagonia, Rio Negro, (2011)

Argentina

Me Neuquen, Neuquen, Rio Colorado 85 -43.85 Calvo et al. (1997)

Patagonia, Argentina

Me Sierra de Los Llanos, La Los Llanos 80 -35.50 Hechenleitner et al.

Rioja, Argentina (2016) 287

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Me Baruru Basin, Marilia Marilia 65 -29.95 Ribeiro (2002)

County, San Paulo State,

Brazil

Me Laguna Umayo, Bagua Fundo El Triunfo 75 -21.10 Vianey-Liaud et al.

Basin, Peru (1997, 2003)

Pre Casal da Rola, near Lourinha 150 28.60 Ribeiro et al. (2014)

Lourinha, Portugal

Pre Pai Mogo, Lourinha, Lourihna 150 28.40 Mateus et al. (1997);

Portugal Antunes et al. (1998)

Pre Freemont County, Colorado, Morrison 155.7–145.5 34.00 Zelenitsky (2004)

USA

Pre Delta and Montrose County, Morrison 155.7–150.8 33.70 Hirsch (1994)

Colorado, USA

Pre Cleveland-Lloyd Dinosaur Morrison 155.7–150.8 34.90 Hirsch et al. (1989);

Quarry, Emery County, Hirsch (1994) 288

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Utah, USA

Pre Fox Mesa, Wyoming, USA Morrison 155.7–145.5 40.30 Carrano et al. (2013)

Pri Oukdiksou syncline, Irbzer 65 19.95 Garcia et al. (2003)

Morocco

Pri Heyuan Basin, Guangdong, Dongyuan 65 29.10 Lü et al. (2006);

China Tanaka et al. (2012)

Pri Xixia Basin (or Majiacun (or 85 37.85 Liang et al. (2009)

Xinping-Chimei Basin), Zhaoying)

Xixia, Henan, China

Pri Xixia Basin (or Gaogou 90 38.60 Wang and Zhou

Xinping-Chimei Basin), (1995)

Xixia, Henan, China

Pri Hejiagou, Yunxian, Hubei, Gaogou 90 37.70 Zhou et al. (1998)

China

Pri Bayan Mandahu, Urad Djadokhta 70 44.70 Zhao and Li (1993) 289

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Houqi, Inner Mongolia,

China

Pri Ganzhou, Jiangxi, China Nanxiong 65 31.20 Tanaka pers. obs. at

Ganzhou Museum,

Jiangxi Province,

China

Pri Tiantai Basin, Zhejiang, Chichengshan and 95 36.35 Fang et al. (2000,

China Laijia 2003); Wang et al.

(2011)

Pri Tiantai Basin, Zhejiang, Liangtoutang 100 36.90 Varricchio et al.

China (2015)

Pri Kamitaki, Tamba, Hyogo, 'Lower Formation' of 110 49.00 Tanaka et al. (2016)

Japan Sasayama Group

Pri Khermin-Tsav, Gobi, Barun-Goyot 75 45.25 Mikhailov (1994b)

Mongolia 290

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Pri Baga-Tariach, Gobi, Dzhadokhta 70 48.90 Mikhailov (1994b)

Mongolia

Pri Tugrikin Shire, Gobi, Djadokhta 70 46.50 Watabe et al. (2010)

Mongolia

Pri Zhinst Tolgoi, Udyn Sayr, Djadokhta 70 46.40 Suzuki and Watabe

Gobi, Mongolia (2000a)

Pri Bayn Dzak, Gobi, Mongolia Djadokhta 70 46.50 Watabe et al. (2010)

Pri Khashaat, Gobi, Mongolia Dzhadokhta 70 46.50 Sabath (1991)

Pri Altan Ula IV, Gobi, Nemegt 70 45.50 Sabath (1991)

Mongolia

Pri Kakanaut, Koryak Upland, Kakanaut 65 76.70 Godefroit et al. (2009)

Chukotka Autonomous

Region, Russia

Pri Herault region, France Villeveyrac-Meze 75 32.85 Garcia et al. (2000)

Basin 291

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Pri Aix-en-Provence region, NA 75 32.95 Garcia et al. (2000)

France

Pri Teruel, Spain Belsa 130 25.60 Canudo et al. (2010);

Moreno-Azanza et al.

(2014b)

Pri Southern Pyrenees region Tremp 65 29.65 Vianey-Liaud and

(e.g., Coll de Nargo and Lopez-Martinez

Tremp), Spain (1997)

Pri Southern Pyrenees region Aren 70 30.00 Lopez-Martinez and

(e.g., Coll de Nargo and Vicens (2012)

Tremp), Spain

Pri Dinosaur Provincial Park, Dinosaur Park 75 57.05 Zelenitsky and

Alberta Sloboda (2005)

Pri Devil's Coulee, Alberta, Oldman 75 56.05 Zelenitsky and Hills

Canada (1996) 292

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Pri Near Choteau, Teton Two Medicine 75 54.55 Varricchio et al.

County, Montana; Glacier (2002)

County, Montana, USA

Pri First Find Microsite, Two Medicine 80 54.00 Jackson and

Sevenmile Hill outcrops, Varrichhio (2010)

Teton County, Montana,

USA

Pri Hill County, Montana, USA Judith River 75 54.75 Jackson et al. (2010)

Pri Bone Hill Locality, Red Two Medicine 85 53.10 Hirsch and Quinn

Rock Locality, Teton (1990)

County, Montana; Glacier

County, Montana, USA

Pri NMMNH L-4010, Fruitland 84.9–70.6 43.60 Tanaka et al. (2011)

Bisti/De-na-zin Wilderness

area, San Juan Basin, New 293

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Mexico, USA

Pri Lizard and Sauropod North Horn 70.6–66.0 47.20 Bray (1999)

localities, Emery County,

Utah, USA

Pri UCM L79059, Mudbank North Horn 70.6–66.0 46.10 Bray (1999)

Site 1.2, Emery County,

Utah, USA

Pri El Pantano, Coahuila, Cerro del Pueblo 84.9–70.6 31.80 Aguillon-Martinez et

Mexico al. (2004)

Sp Heyuan Bsin, Guangdong, Nanxiong 65 29.10 Zhang and Huang

China (1999)

Sp Maoming Basin, Tongguling 80 25.65 Zhang and Huang

Guangdong, China (1999)

Sp Nanxiong Basin (or Nanxiong and 65 30.35 Zhang and Huang

Nanxiong-Shixing Basin), Pingling (1999) 294

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Guangdong, China

Sp Sanshui-Guangzhou Basin, Sanshui 90 28.80 Zhang and Huang

Guangdong, China (1999)

Sp Sanshui-Guangzhou Basin, Baihedong Group 125 26.25 Zhang and Huang

Guangdong, China (1999)

Sp Loushan Basin, Henan, Majiacun (or 85 37.25 Liang et al. (2009)

China Zhaoying)

Sp Wulichuan Basin, Henan, Majiacun (or 85 38.35 Liang et al. (2009)

China Zhaoying)

Sp Wulichuan Basin, Henan, Gaogou (or 90 39.00 Liang et al. (2009)

China Zoumagang)

Sp Xiaguan Basin, Henan, Majiacun (or 85 37.90 Liang et al. (2009)

China Zhaoying)

Sp Xichuan Basin, Henan, Majiacun (or 85 37.70 Liang et al. (2009)

China Zhaoying) 295

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Sp Xichuan Basin, Henan, Gaogou (or 90 38.40 Liang et al. (2009)

China Zoumagang)

Sp Xixia Basin (or Majiacun (or 85 37.85 Liang et al. (2009)

Xinping-Chimei Basin), Zhaoying)

Xixia, Henan, China

Sp Xixia Basin (or Gaogou (or 90 38.60 Liang et al. (2009)

Xinping-Chimei Basin), Zoumagang)

Xixia, Henan, China

Sp Xixia Basin (or Sigou (or Luyemiao 75 37.25 Liang et al. (2009)

Xinping-Chimei Basin), or Hugang)

Xixia, Henan, China

Sp Yangji Basin, Henan, China Majiacun (or 85 38.10 Liang et al. (2009)

Zhaoying)

Sp Qinglongshan, Yunxian, Gaogou 90 37.70 Zhou et al. (1998)

Hubei, China 296

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Sp Changtu, Liaoning, China Quantou 110 45.40 Liu et al. (2013)

Sp Laiyang-Zhucheng, Jiangjunding 85 42.85 Young (1965); Zhao

Shangdong, China Formation in the and Jiang (1974)

Wangshi Group

Sp Laiyang-Zhucheng, Jingangkou 75 41.25 Zhao et al. (2013)

Shangdong, China Formation in the

Wangshi Group

Sp Tiantai Basin, Zhejiang, Chichengshan and 95 36.35 Fang et al. (2003,

China Laijia 2000); Wang et al.

(2011)

Sp Kholdoda and Chandrapur, Lameta 65 -22.45 Mohabey (1996,

Maharashtra, India 2001)

Sp Kamitaki, Tamba, Hyogo, 'Lower Formation' of 110 49.00 Tanaka et al. (2016)

Japan Sasayama Group

Sp Shiluut Ula, Gobi, Mongolia Barungoyot 75 45.20 Watabe et al. (2010) 297

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Sp Abdrant Nuru, Gobi, ?Djadokhta 70 46.60 Watabe (2004)

Mongolia

Sp Baga-Tariach, Gobi, Djadokhta 70 48.90 Mikhailov (1994a)

Mongolia

Sp Bortolgoi, Gobi, Mongolia Djadokhta 70 46.20 Watabe et al. (2010)

Sp Dzamin Khond, Gobi, Djadokhta 70 46.30 Watabe et al. (2010)

Mongolia

Sp Dragon's Tomb' dinosaur Nemegt 70 45.40 Dewaele et al. (2015)

locality, Gobi, Mongolia

Sp Shar Tsav, Gobi, Mongolia Nemegt 70 46.80 Watabe et al. (2010)

Sp Yagaan Khovil, Gobi, ?Nemegt 70 46.20 Watabe et al. (2010)

Mongolia

Sp Kakanaut, Koryak Upland, Kakanaut 65 76.70 Godefroit et al. (2009)

Chukotka Autonomous

Region, Russia 298

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Sp Bosung, Chullanam-do, Seonso 80 41.00 Huh and Zelenitsky

South Korea (2002)

Sp Goseong, South Korea Goseong 75 41.60 Paik et al. (2012)

Sp Teruel, Spain Blesa 130 25.60 Moreno-Azanza et al.

(2014a)

Sp Teruel, Spain Camarillas, 130 25.50 Moreno-Azanza et al.

Marambel, and Upper (2014a)

El Castellar

Sp Southern Pyrenees region Tremp 65 29.65 Sellés et al. (2014)

(e.g., Coll de Nargo and

Tremp), Spain

Sp Dinosaur Provincial Park, Dinosaur Park and 75 57.05 Zelenitsky and

Alberta Oldman Sloboda (2005)

Sp Devil's Coulee, Alberta, Oldman 75 56.05 Zelenitsky and Hills

Canada (1997) 299

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Sp Knight's Ranch, Alberta, Oldman 75 55.95 Zelenitsky and Hills

Canada (1997)

Sp Near Choteau, Teton Two Medicine 75 54.55 Horner and Makela

County, Montana, USA (1979)

Sp Sevenmile Hill, Teton Two Medicine 80 54.00 Jackson and

County, Montana, USA Varrichhio (2010)

Sp Blacktail Creek, Glacier Two Medicine 75 55.35 Horner and Currie

County, Montana, USA (1994)

Sp One Place Locality, Red Two Medicine 85 53.10 Hirsch and Quinn

Rock Locality, Teton (1990)

County, Montana; Glacier

County, Montana

Sp Lizard Locality, Emery North Horn 70.6–66.0 47.20 Bray (1999)

County, Utah, USA

Sp El Pantano, Coahuila, Cerro del Pueblo 84.9–70.6 31.80 Aguillon-Martinez et 300

Ootaxon Locality Formation Approximate Approximate Example of references

age (Ma) paleolatitude (°)

Mexico al. (2004)

301

Appendix 5.1. Comparison of Ip values measured from naturally and artificially incubated eggs of extant crocodylians and megapodes.

In order to evaluate the agreement between two methods, in this case, values of Ip measured from naturally and artificially incubated eggs of crocodylian and megapode species, a Bland–Altman plot (Bland and Altman, 1986) and Passing–Bablok linear regression analysis (Passing and Bablok, 1983) were conducted using the dataset of Appendix 5.6A and following the methodology in Tanaka and Zelenitsky (2014a). The Bland–Altman plot graphically evaluates two methods by comparing the differences of each pair of variables (i.e., log Ip from naturally incubated egg – log Ip from artificially incubated egg) in y-axis with the mean values of two variables [i.e., 0.5·(log Ip from naturally incubated egg + log Ip from artificially incubated egg)] in x-axis. A correlation between differences (y-axis) and means (x-axis) was tested with both phylogenetic and non-phylogenetic approaches because a linear relationship between the variables indicates a proportional error (Bland and Altman, 1995). A correlation by non-phylogenetic approach was implemented with IBM SPSS Statistics v. 22.0.0 (IBM SPSS Inc.). For the phylogenetic approach, the independent contrast method (Felsenstein, 1985) was applied using PDAP module v.1.16 (Midford et al., 2010) of the software Mesquite 3.02 (Maddison and Maddison, 2010). A phylogenetic tree of crocodylians and megapodes was reconstructed based on Crowe et al. (2006), Oaks (2011), and Tanaka et al. (2015) (see "5.2 Material and Methods" for the methodology: Appendix 5.9). In the Bland–Altman plot, the upper and lower 95% limits of agreement (i.e., , where is the mean difference and SD is the standard deviation of the mean difference: Altman and Bland, 1983) were also calculated in order to show how much the two types of Ip measurements agree or differ. Additionally, a Passing–Bablok linear regression analysis was implemented in order to examine systematic errors between log Ip values from naturally and artificially incubated eggs. Constant and proportional errors can be detected when the 95% confidence interval of the intercept does not include zero and when the confidence interval of the slope does not include one, respectively. The Passing–Bablok linear regression was conducted with MedCalc Statistical Software version 16.4.3 (MedCalc Software bvba, Ostend, Belgium: http://www.medcalc.org).

Results indicated that there is agreement between values of Ip measured from naturally and artificially incubated eggs (Appendices 5.10 and 5.11). In the Bland–Altman plot, the mean of the differences was 0.005, suggesting that the mean Ip value of naturally incubated eggs is

302

almost identical to that of artificially incubated eggs. The lower and upper 95% limits of agreement were -0.077 and 0.088, respectively, indicating that for 95% of cases, Ip values of artificially incubated eggs will be only between 0.84 and 1.22 times (i.e., antilog of -0.077 and 0.088, respectively) the values of naturally incubated eggs. Also, the Bland–Altman plot revealed no significant correlation between the differences and the mean values with both conventional and phylogenetically corrected approaches (p = 0.414 for conventional approach and p = 0.466 and 0.794 for phylogenetically corrected approaches: Appendix 5.10). The Passing–Bablok regression did not detect systematic errors (i.e., constant error and proportional error) between the two types of Ip measurements; the 95% confidence intervals of the intercept and the slope include zero (-1.901 to 0.257) and one (0.861 to 2.010), respectively. These results suggest that Ip values measured from naturally incubated eggs are comparable to those from artificially incubated eggs.

303

Appendix 5.2. Specimens examined for this study. Institutional abbreviations: AIM, Auckland Institute and Museum, Auckland, New Zealand; CM, Carnegie Museum of Natural History, Pittsburgh, Pennsylvania; HEC, Hirsch Egg Catalogue, University of Colorado Museum, Boulder, Colorado; MCZ, Museum of Comparative Zoology, Cambridge, Massachusetts; ROM, Royal Ontario Museum, Toronto, Canada; TMP, Royal Tyrrell Museum of Palaeontology, Drumheller, Canada; UHR, Hokkaido University Museum, Sapporo, Japan; YPM R., Herpetology Collection at the Yale Peabody Museum, New Haven, Connecticut; ZEC, Zelenitsky Egg catalogue, University of Calgary, Calgary, Canada. Specimen Taxon Collection number Modern eggshell Aechmophorus occidentalis ROM 4693 Aix galericulata UHR 33131 Aix sponsa UHR 33128 Alectura lathami ZEC 137-1-3; ZEC 137-1-4; ZEC 137-1-5 Alligator mississippiensis YPM HERR. 015109; YPM HERR. 015110 Alligator sinensis YPM HERR. 018989 Ammoperdix heyi ROM 9516 Anas bahamensis ROM 9487 Anas discors ROM 7563 Anas platyrhynchos ROM 12821 Anhinga anhinga ROM 9438 Anser anser ZEC 221-1-2; ZEC 221-1-3 Branta canadensis ZEC 444-1-1; ZEC 444-1-2 Bucephala islandica ROM 7739 Burhinus oedicnemus ROM 10983 Buteo rufinus ROM 8126 Caiman crocodilus UHR 33214 Caiman latirostris UHR 27367; YPM HERR. 017953; YPM HERR. 018990 Caiman yacare UHR 33215; YPM HERR. 019030

304

Specimen Taxon Collection number Cairina moschata ZEC 290-1-1; ZEC 290-1-2 Chlidonias niger ROM 10266 Chrysolophus amherstiae ROM 3659 Clangula hyemalis ROM 12427 Crocodylus moreletii YPM HERR. 018979; YPM HERR. 018980; YPM HERR. 018981 Crocodylus niloticus MCZ 26933; YPM HERR. 017955; YPM HERR. 018982; YPM HERR. 018983; YPM HERR. 018984; YPM HERR. 018985; YPM HERR. 018986; ZEC 136 (HEC175) Crocodylus porosus UHR 33210; YPM R17954; YPM R17976 Crocodylus rhombifer YPM HERR. 011637; YPM HERR. 015102 Crocodylus siamensis YPM HERR. 018977; YPM HERR. 018978 Egretta thula ROM 4772 Egretta tricolor ROM 4782 Eudocimus albus ROM 4818 Eudromia elegans ZEC 283-1-2 Falco naumanni ROM 5096 Falco tinnunculus ROM 10855 Fratercula arctica ROM 2354 Fulmarus glacialis ROM 3034 Gavia immer ROM 13041 Gavialis gangeticus YPM HERR. 018824 Larus glaucescens ROM 5503

305

Specimen Taxon Collection number Larus heermanni ROM 8686 Larus ridibundus ROM 3780 Leipoa ocellata ZEC 218-2-1; ZEC218-2-2 Lophura nycthemera ROM 4622 Megapodius decollatus YPM 142019 Melanosushus niger CM41452 (ZEC303); MCZ 46554 Nycticorax nycticorax UHR 33121 Oceanodroma leucorhoa ROM 2796 Onychoprion fuscatus ROM 5632 Osteolaemus tetraspis YPM HERR. 018823; YPM HERR. 018988 Oxyura jamaicensis ROM 8088 Paleosuchus palpebrosus CM41453 (ZEC304); YPM HERR. 017952 Paleosuchus trigonatus YPM HERR. 018987 Pavo cristatus UHR 33126 Phalacrocorax pelagicus ZEC 445-1-1; ZEC 445-1-2 Plegadis falcinellus ROM 8002 Podiceps cristatus ROM 1182 Podilymbus podiceps ROM 12592 Pygoscelis adeliae ROM 11421 Rissa tridactyla ROM 356 Rynchops niger ROM 10003 Somateria mollissima ROM 10863 Spheniscus demersus ROM 9973 Sterna hirundo ROM 8478 Sterna paradisaea ROM 5615 Sternula albifrons ROM 10864 Streptopelia turtur ROM 3665 Strix aluco ROM 3767

306

Specimen Taxon Collection number

Syrmaticus soemmerringii UHR 33127 Tadorna tadorna ROM 9527 Thalasseus elegans ROM 13086 Thalasseus maximus ROM 5660 Tomistoma schlegelii YPM HERR. 018975; YPM HERR. 018976 Tyto alba ROM 12631 Fossil/sub-fossil Euryapteryx sp. AIM LB6672; AIM LB6673 eggshell Pachyornis geranoides AIM LB6675 Prismatoolithus levis (Troodon TMP1994.179.1 formosus)

307

Appendix 5.3. Values of incubation period (days) of living archosaur species used for analyses.

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Crocodylia Alligator mississippiensis 63.000 84.000 70.650 1.849 Deeming and Ferguson (1989); Joanen and McNease (1989); Lang and Andrews (1994) Alligator sinensis 70.000 72.000 71.000 1.851 Wink and Elsey (1994) Caiman crocodilus 71.818 90.060 80.293 1.905 Lang and Andrews (1994) Caiman latirostris 64.800 84.600 74.733 1.874 Pina et al. (2003) Caiman yacare 56.700 82.000 71.563 1.855 Brazaitis (1986); Miranda et al. (2002) Crocodylus mindorensis 77.000 89.000 82.500 1.916 Alcala et al. (1987); Sibal et al. (1992) Crocodylus moreletii 70.270 94.380 80.294 1.905 Hunt (1973, 1975); Lang and Andrews (1994) Crocodylus niloticus 84.400 111.000 95.133 1.978 Hutton (1987) Crocodylus rhombifer 78.000 1.892 Birchard and Marcellini (1996) Crocodylus porosus 74.000 112.000 88.380 1.946 Webb et al. (1987); Whitehead and Seymour (1990); Sibal et al. (1992); Wilken et al. (1992) Crocodylus siamensis 67.857 91.200 78.352 1.894 Lang and Andrews (1994) Gavialis gangeticus 59.829 84.000 71.276 1.853 Lang and Andrews (1994) Melanosuchus niger 90.000 1.954 Vieira et al. (2011) Osteolaemus tetraspis 87.000 97.000 92.000 1.964 Beck (1978); McCartney (1990)

308

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Paleosuchus palpebrosus 92.000 1.964 Birchard and Marcellini (1996) Paleosuchus trigonatus 115.000 118.000 116.500 2.066 Jardine (1981) Tomistoma schlegelii 75.000 90.000 82.500 1.916 Groombridge (1982) Accipitridae Buteo rufinus 36.000 1.556 Ar and Rahn (1985) Anseriformes Aix galericulata 28.000 33.000 29.375 1.468 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Kear (2005) Aix sponsa 25.000 37.000 30.638 1.486 Hepp and Bellrose (1995); Kear (2005) Anas bahamensis 25.000 26.000 25.500 1.407 Kear (2005) Anas discors 21.000 27.000 23.500 1.371 Kear (2005) Anas fulvigula 24.000 28.000 25.500 1.407 Hoyt et al. (1979); Moorman and Gray (1994) Anas gracilis 25.000 31.000 28.000 1.447 Marchant and Higgins (1990); Kear (2005) Anas platyrhynchos 23.000 32.000 27.320 1.436 Cramp (1977); Marchant and Higgins (1990); Snow and Perrins (1998); Drilling et al. (2002); Kear (2005) Anser anser 27.000 30.000 27.375 1.437 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Kear (2005) Anser brachyrhynchus 25.000 28.000 26.625 1.425 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Kear (2005)

309

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Anser cygnoides 28.000 1.447 Hoyt et al. (1979) Anser erythropus 25.000 28.000 26.125 1.417 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Kear (2005) Anser fabalis 25.000 30.000 27.438 1.438 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Kear (2005) Branta canadensis 24.000 30.000 27.938 1.446 Cramp (1977); Snow and Perrins (1998); Mowbray et al. (2002); Kear (2005) Branta h. minima 26.500 1.423 Hoyt et al. (1979); Mowbray et al. (2002) Branta leucopsis 24.000 28.000 24.375 1.387 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Kear (2005) Branta sandvicensis 29.000 32.000 30.000 1.477 Hoyt et al. (1979); Banko et al. (1999); Kear (2005) Bucephala islandica 26.000 36.000 30.640 1.486 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Eadie et al. (2000); Kear (2005) Cairina moschata 35.000 1.544 Kear (2005) Cereopsis novaehollandiae 34.000 37.000 35.500 1.550 Marchant and Higgins (1990); Kear (2005) Chloephaga melanoptera 30.000 1.477 Kear (2005) Clangula hyemalis 24.000 29.000 25.900 1.413 Cramp (1977); Rahn et al. (1983); Snow and Perrins (1998); Robertson and Savard (2002); Kear (2005)

310

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Cyanochen cyanoptera 30.000 34.000 31.000 1.491 Hoyt et al. (1979); Kear (2005) Cygnus atratus 32.000 48.000 38.811 1.589 Marchant and Higgins (1990); Kear (2005) Cygnus buccinator 32.000 37.000 34.750 1.541 Mitchell (1994); Kear (2005) Dendrocygna arborea 30.000 1.477 Hoyt et al. (1979) Dendrocygna autumnalis 25.000 31.000 28.083 1.448 Hoyt et al. (1979); James and Thompson (2001); Kear (2005) Dendrocygna bicolor 19.000 44.000 26.967 1.431 Hoyt et al. (1979); Hohman et al. (2001); Kear (2005) Lophodytes cucullatus 26.000 41.000 31.978 1.505 Hoyt et al. (1979); Dugger et al. (1994); Kear (2005) Mergus m. merganser 28.000 35.000 31.400 1.497 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Mallory and Metz (1999); Kear (2005) Mergus serrator 28.000 35.000 30.480 1.484 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Titman (1999); Kear (2005) Oxyura jamaicensis 20.000 26.000 24.675 1.392 Cramp (1977); Snow and Perrins (1998); Brua (2001); Kear (2005) Oxyura leucocephala 22.000 24.000 22.500 1.352 Hoyt et al. (1979); Kear (2005) Somateria mollissima 23.000 30.000 26.125 1.417 Cramp (1977); Snow and Perrins (1998); Goudie et al. (2000); Kear (2005)

311

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Tadorna cana 30.000 1.477 Kear (2005) Tadorna tadorna 29.000 31.000 29.500 1.470 Cramp (1977); Hoyt et al. (1979); Snow and Perrins (1998); Kear (2005) Tadorna variegata 30.000 35.000 31.667 1.501 Hoyt et al. (1979); Marchant and Higgins (1990); Kear (2005) Charadriiformes Alca torda 25.000 39.000 34.889 1.543 Cramp (1985); Gaston and Jones (1998); Hipfner et al. (2002) Brachyramphus marmoratus 27.000 30.000 28.833 1.460 Nelson (1997); Gaston and Jones (1998); Zimmerman and Hipfner (2007) Burhinus oedicnemus 24.000 27.000 26.333 1.421 Cramp (1983); Ar and Rahn (1985); Snow and Perrins (1998) Cepphus columba 25.300 38.100 29.017 1.463 Ewins (1993); Gaston and Jones (1998); Zimmerman and Hipfner (2007) Cerorhinca monocerata 39.000 52.000 45.167 1.655 Gaston and Dechesne (1996); Gaston and Jones (1998); Zimmerman and Hipfner (2007) Chlidonias niger 20.900 23.300 21.514 1.333 Cramp (1985); Davis and Ackerman (1985); Dunn and Agro (1995); Snow and Perrins (1998) Fratercula arctica 36.000 45.000 40.075 1.603 Cramp (1985); Gaston and Jones (1998);

312

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Snow and Perrins (1998); Lowther et al. (2002) Fratercula cirrhata 41.000 54.000 44.505 1.648 Gaston and Jones (1998); Piatt and Kikaysky (2002b); Zimmermann and Hipfner (2007) Haematopus ostralegus 24.000 27.000 25.500 1.407 Cramp (1983); Snow and Perrins (1998) Larus argentatus 26.000 32.000 28.754 1.459 Cramp (1983); Pierotti and Good (1994); Snow and Perrins (1998) Larus canus 22.500 28.000 25.189 1.401 Cramp (1983); Snow and Perrins (1998); Moskoff and Bevier (2002) Larus fuscus 24.000 27.000 25.500 1.407 Cramp (1983); Snow and Perrins (1998) Larus glaucescens 26.900 28.400 27.233 1.435 Ar and Rahn (1985); Verbeek (1993) Larus heermanni 23.000 1.362 Ar and Rahn (1985) Larus marinus 26.000 28.000 27.500 1.439 Cramp (1983); Good (1998); Snow and Perrins (1998) Larus ridibundus 23.000 26.000 23.800 1.377 Cramp (1983); Ar and Rahn (1985); Snow and Perrins (1998) Numenius phaeopus 22.000 28.000 26.900 1.430 Cramp (1983); Ar and Rahn (1985); Skeel and Mallory (1996); Snow and Perrins (1998) Onychoprion fuscatus 26.000 31.000 28.267 1.451 Ar and Rahn (1985); Snow and Perrins

313

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References (1998); Schreiber et al. (2002) Pluvialis apricaria 27.000 34.000 30.400 1.483 Cramp (1983) Ptychoramphus aleuticus 37.000 57.000 38.267 1.583 Manuwal and Thoresen (1993); Gaston and Jones (1998); Zimmermann and Hipfner (2007) Rissa tridactyla 25.000 32.000 27.025 1.432 Maunder and Threlfall (1972); Cramp (1983); Ar and Rahn (1985); Snow and Perrins (1998) Rynchops niger 21.000 25.000 22.578 1.354 Grant et al. (1984); Ar and Rahn (1985); Gochfeld and Burger (1994) Stercorarius skua 26.000 32.000 29.361 1.468 Cramp (1983); Higgins and Davies (1996); Snow and Perrins (1998) Sterna hirundo 21.000 23.100 21.592 1.334 Ar and Rahn (1985); Cramp (1985); Snow and Perrins (1998); Nisbet (2002) Sterna paradisaea 20.000 34.000 22.088 1.344 Ar and Rahn (1985); Cramp (1985); Snow and Perrins (1998); Hatch (2002) Sternula albifrons 18.000 25.000 20.917 1.320 Ar and Rahn (1985); Cramp (1985); Snow and Perrins (1998) Synthliboramphus antiquus 28.000 47.000 32.600 1.513 Gaston (1994); Gaston and Jones (1998); Zimmerman and Hipfner (2007) Thalasseus elegans 23.000 33.000 26.650 1.426 Burness et al. (1999)

314

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Thalasseus maximus 28.000 35.000 29.700 1.473 Cramp (1985); Snow and Perrins (1998); Buckley and Buckley (2002) Uria aalge 26.000 49.000 33.273 1.522 Cramp (1985); Gaston and Jones (1998); Snow and Perrins (1998); Ainley et al. (2002); Zimmermann and Hipfner (2007) Ciconiiformes Egretta thula 21.880 24.900 22.733 1.357 Ar and Rahn (1985); Parsons and Master (2000); Kushlan and Hancock (2005) Egretta tricolor 21.000 24.000 22.844 1.359 Ar and Rahn (1985); Frederick (1997); Kushlan and Hancock (2005) Eudocimus albus 21.750 1.337 Ar and Rahn (1985); Kushlan and Bildstein (1992) Nycticorax nycticorax 21.000 26.000 22.760 1.357 Cramp (1977); Ar and Rahn (1985); Davis (1993); Snow and Perrins (1998); Kushlan and Hancock (2005) Plegadis falcinellus 21.500 1.332 Cramp (1977); Ar and Rahn (1985); Snow and Perrins (1998); Davis and Kricher (2000) Columbiformes Columba livia 16.000 19.000 17.800 1.250 Cramp (1985); Johnston (1992); Higgins and Davies (1996); Snow and Perrins

315

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References (1998) Streptopelia turtur 13.000 16.000 13.667 1.136 Ar and Rahn (1985); Cramp (1985); Snow and Perrins (1998) Falconidae Falco naumanni 28.000 29.000 28.333 1.452 Cramp (1980); Ar and Rahn (1985); Snow and Perrins (1998) Falco tinnunculus 27.000 29.000 28.000 1.447 Cramp (1980); Snow and Perrins (1998) Galliformes Alectura lathami 47.000 48.000 47.500 1.677 Baltin (1969) Ammoperdix heyi 21.000 24.000 20.750 1.317 Ar and Rahn (1985); Snow and Perrins (1998) Chrysolophus amherstiae 22.000 23.000 23.000 1.362 Cramp (1980); Ar and Rahn (1985); Snow and Perrins (1998) Coturnix coturnix 17.000 20.000 18.500 1.267 Cramp (1980); Snow and Perrins (1998) Gallus gallus 18.000 20.000 21.000 1.322 Ar and Rahn (1985) Leipoa ocellata 60.000 70.000 65.000 1.813 Booth (1987) Lophophorus impejanus 29.000 1.462 Ar and Rahn (1985) Lophura nycthemera 25.000 26.000 26.000 1.415 Ar and Rahn (1985) Meleagris gallopavo 25.200 31.000 27.400 1.438 Eaton (1992) Numida meleagris galeatus 24.000 25.000 24.500 1.389 Cramp (1980); Snow and Perrins (1998) Pavo cristatus 27.000 30.000 28.167 1.450 Ar and Rahn (1985); Marchant and Higgins

316

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References (1994); Kannan and James (1998) Phasianus colchicus 23.000 28.000 24.375 1.387 Cramp (1980); Marchant and Higgins (1994); Snow and Perrins (1998); Giudice and Ratti (2001) Syrmaticus soemmerringii 24.000 25.000 25.000 1.398 Ar and Rahn (1985) Gaviiformes Gavia immer 24.000 30.600 25.583 1.408 Cramp (1977); McIntyre and Barr (1997); Snow and Perrins (1998) Passeriformes Menura novaehollandiae 40.000 57.000 47.750 1.679 Lill (1987); Higgins et al. (2001) Turdus merula 10.000 19.000 13.208 1.121 Cramp (1988); Snow and Perrins (1998); Higgins et al. (2006) Pelecaniformes Anhinga anhinga 26.000 30.000 27.800 1.444 Frederick and Siegel-Causey (2000); Nelson (2006) Morus bassanus 42.000 46.000 43.866 1.642 Cramp (1977); Snow and Perrins (1998); Mowbray (2002); Nelson (2006) Phalacrocorax auritus 25.000 29.900 27.633 1.441 Ar and Rahn (1985); Hatch and Weseloh (1999); Nelson (2006) Phalacrocorax carbo 27.000 34.000 29.280 1.467 Cramp (1977); Marchant and Higgins (1990); Snow and Perrins (1998); Hatch et

317

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References al. (2000); Nelson (2006) Phalacrocorax pelagicus 25.000 37.000 30.583 1.485 Ar and Rahn (1985); Hobson (1997); Nelson (2006) Podicipediformes Aechmophorus occidentalis 22.000 24.000 23.333 1.368 Ar and Rahn (1985); Storer and Nuechterlein (1992); Fjeldsa (2004) Podiceps cristatus 25.000 29.000 27.433 1.438 Cramp (1977); Snow and Perrins (1998); Fjeldsa (2004) Podilymbus podiceps 20.500 27.000 23.356 1.368 Davis et al. (1984); Muller and Storer (1999); Fjeldsa (2004) Procellariiformes Diomedea exulans 68.000 83.000 77.200 1.888 Marchant and Higgins (1990); Brooke (2004) Fulmarus glacialis 41.000 57.000 49.779 1.697 Cramp (1977); Hatch and Nettleship (1998); Snow and Perrins (1998); Brooke (2004) Oceanodroma leucorhoa 37.000 50.000 42.170 1.625 Cramp (1977); Rahn and Huntington (1988); Huntington et al. (1996); Snow and Perrins (1998); Brooke (2004) Puffinus pacificus 48.000 56.000 52.737 1.722 Marchant and Higgins (1990); Whittow (1997); Brooke (2004)

318

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Puffinus puffinus 47.000 66.000 51.075 1.708 Cramp (1977); Lee and Haney (1996); Snow and Perrins (1998); Brooke (2004) Puffinus tenuirostris 52.000 55.000 52.667 1.722 Fitzherbert (1985); Marchant and Higgins (1990); Brooke (2004) Sphenisciformes Aptenodytes forsteri 62.000 67.000 64.400 1.809 Marchant and Higgins (1990); Williams (1995) Aptenodytes patagonicus 51.700 57.000 53.483 1.728 Handrich (1989); Marchant and Higgins (1990); Williams (1995) Eudyptes robustus 31.000 37.000 33.350 1.523 Marchant and Higgins (1990); Williams (1995) Pygoscelis adeliae 30.000 43.000 34.973 1.544 Rahn and Hammel (1982); Marchant and Higgins (1990); Williams (1995) Spheniscus demersus 36.000 39.000 37.600 1.575 Williams (1995) Strigiformes Strix aluco 28.000 30.000 29.000 1.462 Ar and Rahn (1985); Cramp (1985); Snow and Perrins (1998) Tyto alba 29.000 35.000 31.515 1.499 Ar and Rahn (1985); Cramp (1985); Marti (1992); Snow and Perrins (1998); Higgins (1999) Struthioniformes

319

Taxon Minimum Ip Maximum Ip Mean Ip log mean Ip References Apteryx australis 85.000 96.000 88.500 1.947 Marchant and Higgins (1990); Davies (2002) Tinamiformes Eudromia elegans 20.000 21.000 20.250 1.306 Ar and Rahn (1985); Davies (2002)

320

Appendix 5.4. Dependent variables used for estimation of incubation period in living archosaur species. −1 Taxon log M log Ap∙Ls log D Nest type References Crocodylia Alligator mississippiensis 1.905 1.465 -0.885 0 Ferguson (1985); Wink et al. (1990); Thorbjarnarson (1996); Marzola et al. (2015); Tanaka et al. (2015); this study Alligator sinensis 1.674 1.121 -0.821 0 Ferguson (1985); Wink and Elsey (1994); Thorbjarnarson et al. (2001); Tanaka et al. (2015); this study Caiman crocodilus 1.785 1.175 -1.217 0 Ferguson (1985); Thorbjarnarson (1996); Tanaka et al. (2015); this study Caiman latirostris 1.880 1.766 -0.705 0 Ferguson (1985); Thorbjarnarson (1996); Stoker et al. (2013); Tanaka et al. (2015); this study Caiman yacare 1.821 1.253 -0.711 0 Ferguson (1985); Brazaitis (1986); Miranda et al. (2002); Tanaka et al. (2015); this study Crocodylus mindorensis 1.864 1.503 -0.678 0 Ferguson (1985); Alcala et al. (1987); Sibal et al. (1992); Thorbjarnarson (1996); Marzola et al. (2015); Tanaka et al. (2015) Crocodylus moreletii 1.871 1.130 -1.198 0 Thorbjarnarson (1996); Platt et al.

321

−1 Taxon log M log Ap∙Ls log D Nest type References (2008); Tanaka et al. (2015); this study Crocodylus niloticus 2.036 1.300 -1.149 0 Ferguson (1985); Thorbjarnarson (1996); Tanaka et al. (2015); this study Crocodylus rhombifer 2.018 1.325 -1.127 0 Ferguson (1985); Thorbjarnarson (1996); Tanaka et al. (2015); this study Crocodylus porosus 2.035 1.294 -0.847 0 Ferguson (1985); Sibal et al. (1992); Thorbjarnarson (1996); Tanaka et al. (2015); this study Crocodylus siamensis 2.029 1.245 -0.767 0 Ferguson (1985); Thorbjarnarson (1996); Tanaka et al. (2015); this study Gavialis gangeticus 2.208 2.080 -0.470 0 Ferguson (1985); Thorbjarnarson (1996); Tanaka et al. (2015); this study Melanosuchus niger 2.157 1.548 -0.895 0 Ferguson (1985); Thorbjarnarson (1996); Tanaka et al. (2015); this study Osteolaemus tetraspis 1.723 0.783 -1.113 0 Beck (1978); Ferguson (1985); McCartney (1990); Thorbjarnarson (1996); Tanaka et al. (2015); this study Paleosuchus palpebrosus 1.837 1.522 -0.755 0 Ferguson (1985); Thorbjarnarson (1996); Marzola et al. (2015); Tanaka et al. (2015); this study Paleosuchus trigonatus 1.843 1.855 -0.807 0 Ferguson (1985); Rivas et al. (2001);

322

−1 Taxon log M log Ap∙Ls log D Nest type References Tanaka et al. (2015); this study Tomistoma schlegelii 2.146 1.621 -0.872 0 Groombridge (1982); Ferguson (1985); Thorbjarnarson (1996); Tanaka et al. (2015); this study Accipitridae Buteo rufinus 1.783 0.012 -0.379 1 Ar and Rahn (1985); Tanaka et al. (2015); this study Anseriformes Aix galericulata 1.612 0.583 0.009 1 Tullett (1976); Cramp (1977); Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015); this study Aix sponsa 1.625 0.707 0.035 1 Ar and Rahn (1985); Hepp and Bellrose (1995); Kear (2005); Tanaka et al. (2015); this study Anas bahamensis 1.567 0.344 0.158 1 Tullett (1976); Ar and Rahn (1985); Kear (2005); Tanaka et al. (2015); this study Anas discors 1.412 -0.259 -0.092 1 Kear (2005); Tanaka et al. (2015); this study Anas fulvigula 1.737 1.051 0.131 1 Hoyt et al. (1979); Moorman and Gray (1994); Kear (2005); Tanaka et al. (2015) Anas gracilis 1.525 0.527 0.036 1 Hoyt et al. (1979); Kear (2005); Tanaka

323

−1 Taxon log M log Ap∙Ls log D Nest type References et al. (2015) Anas platyrhynchos 1.793 0.520 0.095 1 Cramp (1977); Rokitka and Rahn (1987); Marchant and Higgins (1990); Drilling et al. (2002); Kear (2005); Balkan et al. (2006); Tanaka et al. (2015); this study Anser anser 2.188 0.857 0.008 1 Tullett (1976); Cramp (1977); Hoyt et al. (1979); Rokitka and Rahn (1987); Kear (2005); Tanaka et al. (2015); this study Anser brachyrhynchus 2.121 0.817 -0.007 1 Cramp (1977); Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015) Anser cygnoides 2.159 1.264 0.190 1 Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015) Anser erythropus 2.017 0.902 0.171 1 Cramp (1977); Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015) Anser fabalis 2.172 0.808 -0.031 1 Tullett (1976); Cramp (1977); Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015) Branta canadensis 2.235 1.149 -0.058 1 Tullett (1976); Cramp (1977); Marchant and Higgins (1990); Mowbray et al. (2002); Kear (2005); Tanaka et al. (2015); this study Branta h. minima 2.000 0.383 0.040 1 Hoyt et al. (1979); Mowbray et al.

324

−1 Taxon log M log Ap∙Ls log D Nest type References (2002); Kear (2005); Tanaka et al. (2015) Branta leucopsis 2.022 0.649 -0.131 1 Tullett (1976); Cramp (1977); Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015) Branta sandvicensis 2.189 0.934 -0.042 1 Tullett (1976); Hoyt et al. (1979); Banko et al. (1999); Kear (2005); Tanaka et al. (2015) Bucephala islandica 1.825 0.832 -0.087 1 Cramp (1977); Hoyt et al. (1979); Eadie et al. (2000); Kear (2005); Tanaka et al. (2015); this study Cairina moschata 1.852 0.763 -0.051 1 Tullett (1976); Kear (2005); Tanaka et al. (2015); this study Cereopsis novaehollandiae 2.102 0.427 -0.066 1 Tullett (1976); Marchant and Higgins (1990); Kear (2005); Tanaka et al. (2015) Chloephaga melanoptera 2.085 0.728 -0.127 1 Carey et al. (1990); Kear (2005); Tanaka et al. (2015) Clangula hyemalis 1.624 0.284 0.220 1 Cramp (1977); Rahn et al. (1983); Robertson and Savard (2002); Kear (2005); Tanaka et al. (2015); this study Cyanochen cyanoptera 1.915 0.663 -0.246 1 Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015) Cygnus atratus 2.422 1.264 -0.078 1 Tullett (1976); Marchant and Higgins

325

−1 Taxon log M log Ap∙Ls log D Nest type References (1990); del Hoyo et al. (1992); Kear (2005) Cygnus buccinator 2.523 1.267 -0.131 1 Tullett (1976); Mitchell (1994); Baicich and Harrison (1997); Kear (2005) Dendrocygna arborea 1.785 0.703 0.093 1 Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015) Dendrocygna autumnalis 1.619 0.352 0.171 1 Hoyt et al. (1979); James and Thompson (2001); Kear (2005); Tanaka et al. (2015) Dendrocygna bicolor 1.701 0.570 0.182 1 Hoyt et al. (1979); Hohman et al. (2001); Kear (2005); Tanaka et al. (2015) Lophodytes cucullatus 1.754 0.215 -0.076 1 Tullett (1976); Hoyt et al. (1979); Dugger et al. (1994); Kear (2005); Tanaka et al. (2015) Mergus m. merganser 1.866 0.806 -0.224 1 Cramp (1977); Hoyt et al. (1979); Mallory and Metz (1999); Kear (2005); Tanaka et al. (2015) Mergus serrator 1.847 0.689 0.041 1 Tullett (1976); Cramp (1977); Titman (1999); Kear (2005); Tanaka et al. (2015) Oxyura jamaicensis 1.872 1.145 0.252 1 Tullett (1976); Cramp (1977); del Hoyo et al. (1992); Brua (2001); Kear (2005); this study

326

−1 Taxon log M log Ap∙Ls log D Nest type References Oxyura leucocephala 1.977 1.039 0.254 1 Cramp (1977); Hoyt et al. (1979); del Hoyo et al. (1992); Kear (2005) Somateria mollissima 2.038 0.984 0.424 1 Cramp (1977) ; Goudie et al. (2000); Kear (2005); Tanaka et al. (2015); this study Tadorna cana 1.949 0.978 -0.019 1 Tullett (1976); Kear (2005); Tanaka et al. (2015) Tadorna tadorna 1.899 0.888 -0.016 1 Cramp (1977); Hoyt et al. (1979); Kear (2005); Tanaka et al. (2015); this study Tadorna variegata 1.940 1.096 0.080 1 Tullett (1976); Hoyt et al. (1979); Marchant and Higgins (1990); Kear (2005); Tanaka et al. (2015) Charadriiformes Alca torda 1.970 0.283 -0.134 1 Tullett (1976); Cramp (1985); Gaston and Jones (1998); Hipfner et al. (2002); Tanaka et al. (2015) Brachyramphus marmoratus 1.585 0.442 -0.108 1 Nelson (1997); Gaston and Jones (1998); Zimmermann and Hipfner (2007); Tanaka et al. (2015) Burhinus oedicnemus 1.555 0.677 0.047 1 Cramp (1983); Ar and Rahn (1985); Tanaka et al. (2015); this study

327

−1 Taxon log M log Ap∙Ls log D Nest type References Cepphus columba 1.738 0.524 -0.178 1 Ewins (1993); Gaston and Jones (1998); Zimmermann and Hipfner (2007); Tanaka et al. (2015) Cerorhinca monocerata 1.890 0.224 -0.363 1 Gaston and Dechesne (1996); Gaston and Jones (1998); Zimmermann and Hipfner (2007); Tanaka et al. (2015) Chlidonias niger 1.026 0.339 0.057 1 Harrison (1975); Cramp (1985); Davis and Ackerman (1985); Dunn and Agro (1995); this study Fratercula arctica 1.805 0.696 0.009 1 Tullett (1976); Cramp (1985); Gaston and Jones (1998); Lowther et al. (2002); Tanaka et al. (2015); this study Fratercula cirrhata 1.957 0.677 -0.141 1 Tullett (1976); Gaston and Jones (1998); Piatt and Kikaysky (2002b); Zimmermann and Hipfner (2007); Tanaka et al. (2015) Haematopus ostralegus 1.672 0.301 -0.128 1 Tullett (1976); Cramp (1983); Tanaka et al. (2015) Larus argentatus 1.954 0.664 0.063 1 Tullett (1976); Cramp (1983); Pierotti and Good (1994); Tanaka et al. (2015) Larus canus 1.708 0.577 0.045 1 Tullett (1976); Cramp (1983); Moskoff

328

−1 Taxon log M log Ap∙Ls log D Nest type References and Bevier (2002); Tanaka et al. (2015) Larus fuscus 1.892 0.975 0.049 1 Tullett (1976); Cramp (1983); Tanaka et al. (2015) Larus glaucescens 1.978 0.876 0.062 1 Ar and Rahn (1985); Verbeek (1993); Tanaka et al. (2015); this study Larus heermanni 1.733 0.602 -0.085 1 Rahn and Dawson (1979); Ar and Rahn (1985); Islam (2002); Tanaka et al. (2015); this study Larus marinus 2.057 0.952 -0.018 1 Tullett (1976); Cramp (1983); Good (1998); Tanaka et al. (2015) Larus ridibundus 1.568 0.300 -0.004 1 Cramp (1983); Ar and Rahn (1985); Tanaka et al. (2015); this study Numenius phaeopus 1.728 0.758 0.000 1 Tullett (1976); Cramp (1983); Ar and Rahn (1985); Skeel and Mallory (1996); Tanaka et al. (2015) Onychoprion fuscatus 1.530 0.408 -0.013 1 Ar and Rahn (1985); Schreiber et al. (2002); Tanaka et al. (2015); this study Pluvialis apricaria 1.544 0.289 0.064 1 Tullett (1976); Cramp (1983); Tanaka et al. (2015) Ptychoramphus aleuticus 1.447 -0.141 -0.310 1 Manuwal and Thoresen (1993); Gaston and Jones (1998); Zimmermann and

329

−1 Taxon log M log Ap∙Ls log D Nest type References Hipfner (2007); Tanaka et al. (2015) Rissa tridactyla 1.692 0.384 -0.042 1 Cramp (1983); Ar and Rahn (1985); Baird (1994); Tanaka et al. (2015); this study Rynchops niger 1.425 0.020 0.152 1 Grant et al. (1984); Ar and Rahn (1985); Gochfeld and Burger (1994); Tanaka et al. (2015); this study Stercorarius skua 1.983 0.767 -0.046 1 Tullett (1976); Cramp (1983); Higgins and Davies (1996); Tanaka et al. (2015) Sterna hirundo 1.312 0.284 -0.063 1 Harrison (1975); Ar and Rahn (1985); Cramp (1985); Nisbet (2002); this study Sterna paradisaea 1.274 -0.314 -0.064 1 Ar and Rahn (1985); Cramp (1985); Hatch (2002); Tanaka et al. (2015); this study Sternula albifrons 0.985 -0.396 -0.257 1 Ar and Rahn (1985); Cramp (1985); Tanaka et al. (2015); this study Synthliboramphus antiquus 1.656 0.075 -0.310 1 Gaston (1994); Gaston and Jones (1998); Zimmermann and Hipfner (2007); Tanaka et al. (2015) Thalasseus elegans 1.588 0.265 -0.044 1 Ar and Rahn (1985); Burness et al. (1999); Tanaka et al. (2015); this study

330

−1 Taxon log M log Ap∙Ls log D Nest type References Thalasseus maximus 1.779 0.723 0.061 1 Cramp (1985); Rokitka and Rahn (1987); Buckley and Buckley (2002); Tanaka et al. (2015); this study Uria aalge 2.041 0.709 -0.292 1 Tullett (1976); Cramp (1985); Gaston and Jones (1998); Ainley et al. (2002); Zimmermann and Hipfner (2007); Tanaka et al. (2015) Ciconiiformes Egretta thula 1.354 0.333 0.088 1 Ar and Rahn (1985); Parsons and Master (2000); Tanaka et al. (2015); this study Egretta tricolor 1.367 0.084 0.007 1 Ar and Rahn (1985); Frederick (1997); Tanaka et al. (2015); this study Eudocimus albus 1.701 0.205 -0.026 1 Ar and Rahn (1985); Kushlan and Bildstein (1992); Tanaka et al. (2015); this study Nycticorax nycticorax 1.577 0.147 0.041 1 Cramp (1977); Ar and Rahn (1985); Tanaka et al. (2015); this study Plegadis falcinellus 1.558 0.424 0.009 1 Cramp (1977); Ar and Rahn (1985); Tanaka et al. (2015); this study Columbiformes Columba livia 1.212 0.091 -0.062 1 Tullett (1976); Cramp (1985); Johnston

331

−1 Taxon log M log Ap∙Ls log D Nest type References (1992); Higgins and Davies (1996); Tanaka et al. (2015) Streptopelia turtur 0.911 -0.479 0.046 1 Ar and Rahn (1985); Cramp (1985); Tanaka et al. (2015); this study Falconidae Falco naumanni 1.128 -0.206 -0.276 1 Cramp (1980); Ar and Rahn (1985); Tanaka et al. (2015); this study Falco tinnunculus 1.301 -0.300 -0.222 1 Tullett (1976); Cramp (1980); Tanaka et al. (2015); this study Galliformes Alectura lathami 2.286 1.515 0.467 0 Ar and Rahn (1985); Booth and Thompson (1991); Marchant and Higgins (1994); Jones et al. (1995); Tanaka et al. (2015); this study Ammoperdix heyi 1.154 -0.652 -0.259 1 Ar and Rahn (1985); Tanaka et al. (2015); this study Chrysolophus amherstiae 1.484 0.724 0.154 1 Ar and Rahn (1985); Tanaka et al. (2015); this study Coturnix coturnix 0.903 0.198 0.414 1 Tullett (1976); Cramp (1980); Tanaka et al. (2015) Gallus gallus 1.765 0.273 0.115 1 Schonwetter (1960–1967); Tullett

332

−1 Taxon log M log Ap∙Ls log D Nest type References (1976); Ar and Rahn (1985); Rokitka and Rahn (1987); Tanaka et al. (2015) Leipoa ocellata 2.250 1.680 0.146 0 Tullett (1976); Booth and Seymour (1987); Marchant and Higgins (1994); Jones et al. (1995); Tanaka et al. (2015); this study Lophophorus impejanus 1.816 0.337 0.093 1 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1985); Tanaka et al. (2015) Lophura nycthemera 1.601 0.877 0.237 1 Schonwetter (1960–1967); Tullett (1976); Ar and Rahn (1985); Tanaka et al. (2015); this study Meleagris gallopavo 1.844 0.748 -0.026 1 Tullett (1976); Rokitka and Rahn (1987); Eaton (1992); Tanaka et al. (2015) Numida meleagris galeatus 1.689 0.205 0.121 1 Tullett (1976); Cramp (1980); Ancel and Girard (1992); Tanaka et al. (2015) Pavo cristatus 1.990 1.163 0.221 1 Ar and Rahn (1985); Kannan and James (1998); Tanaka et al. (2015); this study Phasianus colchicus 1.509 0.417 0.223 1 Tullett (1976); Cramp (1980); Marchant and Higgins (1994); Giudice and Ratti (2001); Tanaka et al. (2015)

333

−1 Taxon log M log Ap∙Ls log D Nest type References Syrmaticus soemmerringii 1.498 0.436 0.228 1 Ar and Rahn (1985); Tanaka et al. (2015); this study Gaviiformes Gavia immer 2.156 1.256 0.282 1 Tullett (1976); Cramp (1977); Tullett and Board (1977); Baicich and Harrison (1997); McIntyre and Barr (1997); this study Passeriformes Menura novaehollandiae 1.780 0.044 -0.276 1 Lill (1987); Higgins et al. (2001); Tanaka et al. (2015) Turdus merula 0.851 -0.235 0.275 1 Tullett (1976); Cramp (1988); Higgins et al. (2006); Tanaka et al. (2015) Pelecaniformes Anhinga anhinga 1.556 0.468 -0.051 1 Colacino et al. (1985); Frederick and Siegel-Causey (2000); Nelson (2006); Tanaka et al. (2015); this study Morus bassanus 2.025 0.396 -0.165 1 Tullett (1976); Cramp (1977); Mowbray (2002); Nelson (2006); Tanaka et al. (2015) Phalacrocorax auritus 1.681 0.443 -0.011 1 Tullett (1976); Ar and Rahn (1985); Hatch and Weseloh (1999); Nelson

334

−1 Taxon log M log Ap∙Ls log D Nest type References (2006); Tanaka et al. (2015) Phalacrocorax carbo 1.814 0.352 -0.021 1 Tullett (1976); Cramp (1977); Marchant and Higgins (1990); Hatch et al. (2000); Nelson (2006); Tanaka et al. (2015) Phalacrocorax pelagicus 1.627 0.448 -0.085 1 Ar and Rahn (1985); Hobson (1997); Nelson (2006); Tanaka et al. (2015); this study Podicipediformes Aechmophorus occidentalis 1.693 1.209 0.281 1 Ar and Rahn (1985); Storer and Nuechterlein (1992); Baicich and Harrison (1997); this study Podiceps cristatus 1.623 1.058 0.397 1 Lomholt (1976); Cramp (1977); Fjeldsa (2004); this study Podilymbus podiceps 1.339 0.958 0.403 1 Davis et al. (1984); Muller and Storer (1999); Fjeldsa (2004); this study Procellariiformes Diomedea exulans 2.681 1.364 -0.219 1 Tullett (1976); Marchant and Higgins (1990); Brooke (2004); Tanaka et al. (2015) Fulmarus glacialis 1.992 1.184 -0.287 1 Cramp (1977); Rahn et al. (1984); Hatch and Nettleship (1998); Brooke (2004);

335

−1 Taxon log M log Ap∙Ls log D Nest type References Tanaka et al. (2015); this study Oceanodroma leucorhoa 1.006 -0.307 -0.142 1 Cramp (1977); Rahn and Huntington (1988); Huntington et al. (1996); Brooke (2004); Tanaka et al. (2015); this study Puffinus pacificus 1.778 0.820 -0.018 1 Tullett (1976); Marchant and Higgins (1990); Whittow (1997); Brooke (2004); Tanaka et al. (2015) Puffinus puffinus 1.765 0.595 -0.112 1 Tullett (1976); Cramp (1977); Lee and Haney (1996); Brooke (2004); Tanaka et al. (2015) Puffinus tenuirostris 1.947 0.951 -0.229 1 Fitzherbert (1985); Marchant and Higgins (1990); Brooke (2004); Tanaka et al. (2015) Sphenisciformes Aptenodytes forsteri 2.658 1.199 -0.268 1 Tullett (1976); Marchant and Higgins (1990); Williams (1995); Tanaka et al. (2015) Aptenodytes patagonicus 2.486 1.153 -0.260 1 Tullett (1976); Handrich (1989); Marchant and Higgins (1990); Williams (1995); Tanaka et al. (2015) Eudyptes robustus 2.041 0.817 -0.100 1 Marchant and Higgins (1990); Williams

336

−1 Taxon log M log Ap∙Ls log D Nest type References (1995); Massaro and Davis (2005); Tanaka et al. (2015) Pygoscelis adeliae 2.066 0.899 -0.082 1 Rahn and Hammel (1982); Marchant and Higgins (1990); Williams (1995); Tanaka et al. (2015); this study Spheniscus demersus 2.024 0.725 -0.171 1 Tullett (1976); Williams (1995); Tanaka et al. (2015); this study Strigiformes Strix aluco 1.580 0.207 -0.275 1 Ar and Rahn (1985); Cramp (1985); Tanaka et al. (2015); this study Tyto alba 1.306 0.071 -0.349 1 Ar and Rahn (1985); Cramp (1985); Marti (1992); Tanaka et al. (2015); this study Struthioniformes Apteryx australis 2.637 1.033 -0.401 1 Silyn-Roberts (1983); Marchant and Higgins (1990); Davies (2002); Tanaka et al. (2015) Tinamiformes Eudromia elegans 1.554 0.408 -0.052 1 Ar and Rahn (1985); Tanaka et al. (2015); this study

337

Appendix 5.5. Dataset of extinct archosaur taxa/ootaxa. −1 Family or oofamily Taxon or ootaxon log M log Ap∙Ls log D References (Possible taxonomic affinity) Cairanoolithidae Cairanoolithus dughii 3.468 2.732 3.425 Williams et al. (1984); Garcia and (Ornithischia?) / Vianey-Liaud (2001); Tanaka et al. Fusioolithidae (2015) Cairanoolithus roussetensis 3.430 2.761 1.630 Garcia and Vianey-Liaud (2001); Tanaka et al. (2015) Megaloolithidae (Sauropod)/ Megaloolithus aureliensis 3.705 3.327 2.230 Garcia and Vianey-Liaud (2001); Fusioolithidae Tanaka et al. (2015) Megaloolithus mammilare 3.716 3.055 2.030 Williams et al. (1984); Garcia and Vianey-Liaud (2001); Tanaka et al. (2015) Megaloolithus 3.351 2.602 4.400 Garcia and Vianey-Liaud (2001); microtuberculata Tanaka et al. (2015) Megaloolithus patagonicus 3.107 2.703 0.933 Jackson et al. (2008); Grellet-Tinner (Titanosaurs) et al. (2012); Tanaka et al. (2015) Megaloolithus petralta 3.420 2.788 2.830 Garcia and Vianey-Liaud (2001); Tanaka et al. (2015) Megaloolithus 3.550 2.842 2.720 Garcia and Vianey-Liaud (2001); pseudomamillare Tanaka et al. (2015) Megaloolithus siruguei 3.622 3.327 3.409 Williams et al. (1984);

338

−1 Family or oofamily Taxon or ootaxon log M log Ap∙Ls log D References (Possible taxonomic affinity) Lopez-Martinez et al. (2000); Garcia and Vianey-Liaud (2001); Deeming (2006); Jackson et al. (2008); Tanaka et al. (2015) Megaloolithus cf. siruguei 3.325 3.138 1.349 Grigorescu et al. (1994, 2010); Tanaka et al. (2015) Megaloolithus sp. 3.267 2.548 2.210 Zelenitsky pers obs. (cited in (recrystarized) Deeming, 2006); Tanaka et al. (2015) Megaloolithus sp. 3.430 3.090 6.250 Zelenitsky pers obs. (cited in (non-recrystarized) Deeming, 2006); Tanaka et al. (2015) Undetermined 3.235 3.085 1.750 Williams et al. (1984); Tanaka et al. megaloolithid oospecies 1 (2015) Undetermined 3.081 2.458 2.220 Grellet-Tinner et al. (2012); Tanaka megaloolithid oospecies 2 et al. (2015) Oofamily Indet. (Non-avian Continuoolithus canadensis 2.320 1.785 0.460 Jackson et al. (2015); Tanaka et al. theropod) (2015) Allosauroidea?/ Lourinhanosaurus antunesi 2.799 2.377 0.250 Antunes et al. (1998); Deeming Coelurosauria? (2006); Tanaka et al. (2015)

339

−1 Family or oofamily Taxon or ootaxon log M log Ap∙Ls log D References (Possible taxonomic affinity) Elongatoolithidae Elongatoolithus andrewsi 2.584 1.621 0.310 Zhao (1975); Mou (1992); Tanaka et (Oviraptorosauria) al. (2015) Elongatoolithus elongatus 2.411 1.659 0.831 Zhao et al. (2013); Tanaka et al. (2015) Macroelongatoolithus 3.488 2.415 0.990 Zelenitsky pers obs. (cited in xixiaensis Deeming, 2006); Tanaka et al. (2015) Macroolithus rugustus 2.772 1.642 0.358 Zhao (1975); Mou (1992); Tanaka et al. (2015) Macroolithus yaotunensis 2.911 1.835 0.285 Zhao (1975); Mou (1992); Tanaka et al. (2015); Wiemann et al. (2015) Prismatoolithidae (Non-avian Prismatoolithus levis 2.463 1.213 0.311 Zelenitsky and Hills (1996); maniraptoran) (Troodon formosus) Varricchio et al. (2013); Tanaka et al. (2015); Zelenitsky pers obs. (cited in Deeming, 2006); this study Sankofa pyrenaica 1.788 0.478 0.630 Lopez-Martinez and Vicens (2012); Tanaka et al. (2015)

340

Appendix 5.6. Comparisons of Ip values derived from naturally and artificially incubated eggs in living covered nesters: (A) dataset of Ip values and (B) results of paired-samples T-test. Parametric T-test was used because homogeneity of variances (Levene’s test) and normal distributions (Shapiro-Wilk test) were assumed (p >> 0.05 in both tests). (A)

Taxon Ip from artificially Ip from naturally incubated eggs incubated eggs Galliformes Alectura lathami 47.500 50.154 Leipoa ocellata 65.000 60.595 Macrocephalon maleo 70.000 73.500 Crocodylia Alligator mississippiensis 70.650 66.194 Caiman crocodilus 80.293 82.000 Caiman yacare 71.563 70.000 Crocodylus johnstoni 86.850 93.000 Crocodylus mindorensis 82.500 97.500 Crocodylus niloticus 95.133 84.000 Crocodylus palustris 78.273 77.350 Crocodylus porosus 88.380 80.333 Osteolaemus tetraspis 92.000 110.000

(B) d.f. t p 11 -0.552 0.592

341

Appendix 5.7. Estimated incubation period ± prediction errors (days) of extinct archosaur taxa/ootaxa based on the multiple regression models (OLS4 to OLS7, PGLS7, and PGLS8).

Prediction errors were calculated based on mean percent prediction errors in Table 5.2.

Taxon or ootaxon OLS4 OLS5 OLS6 OLS7 PGLS7 PGLS8

Cairanoolithidae (Ornithischia?)/ Fusioolithidae

Cairanoolithus dughii 41.820 46.092 48.169 69.427 48.242 63.919

±6.717 ±7.248 ±7.492 ±10.029 ±13.829 ±10.582

Cairanoolithus roussetensis 55.596 59.537 63.475 84.038 54.732 72.366

±8.930 ±9.362 ±9.873 ±12.140 ±15.690 ±11.981

Megaloolithidae (Sauropod)/ Fusioolithidae

Megaloolithus aureliensis 61.391 65.145 69.997 90.299 58.272 76.449

±9.861 ±10.244 ±10.888 ±13.045 ±16.705 ±12.657

Megaloolithus mammilare 57.308 61.211 67.408 90.485 56.018 73.777

±9.205 ±9.625 ±10.485 ±13.071 ±16.058 ±12.215

Megaloolithus 36.294 40.571 41.299 61.346 45.128 59.944 microtuberculata

±5.830 ±6.380 ±6.424 ±8.862 ±12.936 ±9.924

Megaloolithus patagonicus 66.832 70.241 71.751 86.429 59.233 78.287

(Titanosaurs)

±10.735 ±11.045 ±11.160 ±12.486 ±16.980 ±12.961

Megaloolithus petralta 45.820 50.040 51.971 72.268 50.313 66.579

±7.360 ±7.868 ±8.084 ±10.440 ±14.423 ±11.023

Megaloolithus 47.461 51.653 55.046 77.015 51.192 67.681

342

Taxon or ootaxon OLS4 OLS5 OLS6 OLS7 PGLS7 PGLS8 pseudomamillare

±7.624 ±8.122 ±8.562 ±11.126 ±14.675 ±11.205

Megaloolithus siruguei 52.484 56.584 58.708 78.041 54.388 71.422

±8.430 ±8.897 ±9.132 ±11.274 ±15.591 ±11.825

Megaloolithus cf. siruguei 68.789 72.138 73.847 88.408 60.875 80.000

±11.050 ±11.343 ±11.486 ±12.771 ±17.451 ±13.245

Megaloolithus sp. 45.842 50.043 51.632 71.358 49.920 66.261

(recrystarized)

±7.364 ±7.869 ±8.031 ±10.308 ±14.310 ±10.970

Megaloolithus sp. 38.370 42.684 42.145 60.260 47.011 62.038

(non-recrystarized)

±6.163 ±6.712 ±6.555 ±8.705 ±13.476 ±10.271

Undetermined megaloolithid 61.232 64.968 64.795 79.288 57.733 75.976 oospecies 1

±9.836 ±10.216 ±10.078 ±11.454 ±16.550 ±12.579

Undetermined megaloolithid 44.226 48.447 48.314 65.982 48.989 65.113 oospecies 2

±7.104 ±7.618 ±7.515 ±9.532 ±14.043 ±10.780

Oofamily Indet. (Non-avian theropod)

Continuoolithus canadensis 61.220 64.822 61.277 71.702 55.256 73.928

±9.834 ±10.193 ±9.531 ±10.358 ±15.840 ±12.240

Allosauroidea?/ Coelurosauria?

343

Taxon or ootaxon OLS4 OLS5 OLS6 OLS7 PGLS7 PGLS8

Lourinhanosaurus antunesi 95.953 97.172 100.205 106.940 68.690 90.963

±15.413 ±15.280 ±15.586 ±15.449 ±19.691 ±15.060

Elongatoolithidae (Oviraptorosauria)

Elongatoolithus andrewsi 66.561 61.003 71.810 59.665 57.014 42.033

±10.692 ±9.592 ±11.169 ±8.619 ±16.344 ±6.959

Elongatoolithus elongatus 46.923 44.557 48.441 42.996 48.947 36.146

±7.537 ±7.006 ±7.535 ±6.211 ±14.031 ±5.984

Macroelongatoolithus 58.605 54.477 70.362 65.516 55.368 40.441 xixiaensis

±9.414 ±8.566 ±10.944 ±9.464 ±15.872 ±6.695

Macroolithus rugustus 63.603 58.562 71.257 61.574 55.922 41.229

±10.216 ±9.208 ±11.084 ±8.895 ±16.031 ±6.826

Macroolithus yaotunensis 74.456 67.491 84.269 70.267 60.328 44.325

±11.960 ±10.612 ±13.107 ±10.151 ±17.294 ±7.339

Prismatoolithidae (Non-avian maniraptoran)

Prismatoolithus levis 56.939 52.981 62.469 54.745 52.502 38.945

(Troodon formosus)

±9.146 ±8.331 ±9.717 ±7.908 ±15.050 ±6.448

Sankofa pyrenaica 33.215 32.601 33.984 32.817 40.407 30.354

±5.335 ±5.126 ±5.286 ±4.741 ±11.583 ±5.025

344

Appendix 5.8. Estimated incubation period ± prediction errors (days) of extinct archosaur taxa/ootaxa based on the multiple regression models (PGLS9 to PGLS14).

Prediction errors were calculated based on mean percent prediction errors in Table 5.2.

Taxon or ootaxon PGLS9 PGLS10 PGLS11 PGLS12 PGLS13 PGLS14

Cairanoolithidae (Ornithischia?)/ Fusioolithidae

Cairanoolithus dughii 79.944 104.313 41.843 53.569 71.818 105.662

±26.725 ±18.104 ±6.702 9.064 ±14.220 ±15.797

Cairanoolithus 84.998 110.733 54.868 69.393 83.712 120.145 roussetensis

±28.414 ±19.218 ±8.788 11.742 ±16.575 ±17.962

Megaloolithidae (Sauropod)/ Fusioolithidae

Megaloolithus aureliensis 93.397 120.745 60.217 75.340 95.018 134.578

±31.222 ±20.956 ±9.644 12.748 ±18.813 ±20.120

Megaloolithus 94.404 122.460 56.436 71.039 94.916 136.147 mammilare

±31.558 ±21.253 ±9.039 12.021 ±18.793 ±20.355

Megaloolithus 73.911 96.681 36.574 47.172 63.057 93.787 microtuberculata

±24.708 ±16.779 ±5.858 7.982 ±12.485 ±14.022

Megaloolithus 78.232 101.991 65.387 82.120 80.157 112.344 patagonicus (Titanosaurs)

±26.152 ±17.701 ±10.472 13.895 ±15.871 ±16.796

Megaloolithus petralta 79.944 104.216 45.639 58.168 73.636 107.234

345

Taxon or ootaxon PGLS9 PGLS10 PGLS11 PGLS12 PGLS13 PGLS14

±26.725 ±18.087 ±7.309 9.843 ±14.580 ±16.032

Megaloolithus 84.992 110.668 47.188 60.017 79.842 116.270 pseudomamillare

±28.412 ±19.207 ±7.558 10.155 ±15.808 ±17.383

Megaloolithus siruguei 86.184 111.530 51.868 65.320 82.641 118.215

±28.810 ±19.356 ±8.307 11.053 ±16.363 ±17.674

Megaloolithus cf. 83.217 107.875 67.137 83.788 86.077 119.898 siruguei

±27.819 ±18.722 ±10.753 14.178 ±17.043 ±17.926

Megaloolithus sp. 76.506 100.049 45.687 58.393 70.310 102.492

(recrystarized)

±25.575 ±17.364 ±7.317 9.881 ±13.921 ±15.323

Megaloolithus sp. 74.213 96.493 38.516 49.279 63.682 93.390

(non-recrystarized)

±24.809 ±16.747 ±6.169 8.338 ±12.609 ±13.962

Undetermined 77.849 101.060 60.104 75.422 77.059 108.178 megaloolithid oospecies 1

±26.024 ±17.539 ±9.626 12.762 ±15.257 ±16.173

Undetermined 70.441 92.277 44.161 56.587 63.274 92.156 megaloolithid oospecies 2

±23.548 ±16.015 ±7.073 9.575 ±12.528 ±13.778

Oofamily Indet. (Non-avian theropod)

346

Taxon or ootaxon PGLS9 PGLS10 PGLS11 PGLS12 PGLS13 PGLS14

Continuoolithus 59.134 78.110 60.284 76.813 57.405 80.834 canadensis

±19.768 ±13.556 ±9.655 12.998 ±11.366 ±12.085

Allosauroidea?/ Coelurosauria?

Lourinhanosaurus 78.167 102.199 92.338 114.674 89.497 121.631 antunesi

±26.130 ±17.737 ±14.789 19.404 ±17.720 ±18.185

Elongatoolithidae (Oviraptorosauria)

Elongatoolithus andrewsi 68.955 51.432 65.308 58.105 70.705 60.157

±23.051 ±8.926 ±10.460 9.832 ±13.999 ±8.994

Elongatoolithus 57.749 43.149 46.814 42.243 51.866 45.091 elongatus

±19.305 ±7.489 ±7.498 7.148 ±10.269 ±6.741

Macroelongatoolithus 91.404 67.430 57.741 51.166 93.289 80.933 xixiaensis

±30.555 ±11.703 ±9.248 8.658 ±18.471 ±12.100

Macroolithus rugustus 73.726 54.959 62.538 55.733 75.466 64.801

±24.646 ±9.538 ±10.016 9.430 ±14.942 ±9.688

Macroolithus yaotunensis 80.391 59.728 72.624 64.151 87.044 73.775

±26.874 ±10.366 ±11.631 10.855 ±17.234 ±11.030

Prismatoolithidae (Non-avian maniraptoran)

Prismatoolithus levis 65.045 48.793 56.343 50.698 63.695 55.223

347

Taxon or ootaxon PGLS9 PGLS10 PGLS11 PGLS12 PGLS13 PGLS14

(Troodon formosus)

±21.744 ±8.468 ±9.024 8.579 ±12.611 ±8.256

Sankofa pyrenaica 44.658 33.933 33.788 31.363 35.637 32.155

±14.929 ±5.889 ±5.412 5.307 ±7.056 ±4.807

348

Appendix 5.9. Cladogram of 12 living crocodylian and megapode species for Bland–Altman plot.

Cladogram based on Crowe et al. (2006), Oaks (2011), and Tanaka et al. (2015).

349

Appendix 5.10. Bland–Altman plot of differences and means of log Ip values measured from naturally and artificially incubated eggs of crocodylians and megapodes.

The solid line (y = 0.05) represents the mean value of the differences, while the broken lines represent the lower (y = -0.077) and upper (y = 0.088) 95% limits of agreement. Distributions of the differences of log Ip values were assumed to be a normal distribution (Shapiro–Wilk test: p

>> 0.05). There was no significant correlation between values of log differences and log means

(p = 0.414 for non-phylogenetic approach and p = 0.466 and 0.794 for phylogenetic approaches using divergence time and Nee's method for the branch length estimation, respectively).

350

Appendix 5.11. Passing–Bablok regression plot for the dataset of incubation period measured from naturally and artificially incubated eggs of living covered nesters.

The coefficient and intercept of the Passing–Bablok linear regression (solid line) are 1.407 and

-0.766, respectively.

351

Appendix 5.12. Cladogram of 132 living archosaur species for phylogenetic multiple regression analyses.

Cladogram based on Tanaka et al. (2015) with additional phylogenies of Fjeldsa (2004), Thomas et al. (2004), Gonzalez et al. (2009), and Jarvis et al. (2014).

352

Appendix 5.13. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from OLS4 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006).

353

Appendix 5.14. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from OLS5 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006).

354

Appendix 5.15. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from OLS6 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006). The regression of extinct maniraptorans (n = 7, r2 =

0.635) was conducted here based on ordinary least-squares.

355

Appendix 5.16. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from OLS7 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006). The regression of extinct maniraptorans (n = 7, r2 =

0.738) was conducted here based on ordinary least-squares.

356

Appendix 5.17. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from PGLS7 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006). The regression of extinct maniraptorans (n = 7, r2 =

0.615) was conducted here based on ordinary least-squares.

357

Appendix 5.18. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from PGLS8 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006). The regression of extinct maniraptorans (n = 7, r2 =

0.583) was conducted here based on ordinary least-squares.

358

Appendix 5.19. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from PGLS9 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006). The regressions of Megaloolithidae (n = 12, r2 =

0.703) and extinct maniraptorans (n = 7, r2 = 0.948) were conducted here based on ordinary least-squares.

359

Appendix 5.20. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from PGLS11 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006).

360

Appendix 5.21. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from PGLS12 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006).

361

Appendix 5.22. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from PGLS13 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006). The regressions of Megaloolithidae (n = 12, r2 =

0.370) and extinct maniraptorans (n = 7, r2 = 0.879) were conducted here based on ordinary least-squares.

362

Appendix 5.23. Bivariate plots of incubation period and egg mass in both living and extinct archosaur taxa/ootaxa (the estimates of dinosaur Ip derived from PGLS14 in Table 5.2).

The regressions of birds overall (n = 1490, r2 = 0.699) and Procellariiformes (n = 75, r2 = 0.785) were drawn based on Deeming et al. (2006). The regressions of Megaloolithidae (n = 12, r2 =

0.461) and extinct maniraptorans (n = 7, r2 = 0.896) were conducted here based on ordinary least-squares.

363

REFERENCES

Ackerman, R.A. and Platter-Rieger, M., 1979. Water loss by Pied-billed (Podilymbus

podiceps) eggs. American Zoologist 19: 921.

Ackerman, R.A., Whittow, G.C., Paganelli, C.V. and Pettit, T.N., 1980. consumption, gas

exchange, and growth of embryonic Wedge-tailed Shearwaters (Puffinus pacificus

chlororhynchus). Physiological Zoology 53: 210-221.

Agnolin, F.L., Powell, J.E., Novas, F.E. and Kundrat, M., 2012. New alvarezsaurid (Dinosauria,

Theropoda) from uppermost Cretaceous of north-western Patagonia with associated eggs.

Cretaceous Research 35: 33-56.

Aguillon-Martinez, M., Zelenitsky, D.K., Brinkman, D. and Eberth, D., 2004. Eggshell fragments

from the Uppermost Cerro del Pueblo Formation (Upper Cretaceous; Mexico). Journal of

Vertebrate Paleontology 24: 33A.

Ainley, D.G., Nettleship, D.N., Carter, H.R. and Storey, A.E., 2002. (Uria aalge).

In: A. Poole and F. Gill (Editors), The Birds of North America, No. 666. The Birds of North

America, Inc., Philadelphia, pp. 1-44.

Alcala, A.C., Ross, C.A. and Alcala, E.L., 1987. Observations on reproduction and behavior of

captive Philippine Crocodiles (Crocodylus mindorensis Schmidt). Silliman Journal 34:

18-28.

Allsteadt, J., 1994. Nesting ecology of Caiman crocodilus in Cano Negro, Costa Rica. Journal of

Herpetology 28: 12-19.

Altman, D.G. and Bland, J.M., 1983. Measurement in Medicine: the analysis of method

comparison studies. Statistician 32: 307-317.

Ancel, A. and Girard, H., 1992. Eggshell of the domestic Guinea Fowl. British Science 33: 364

993-1001.

Andersen, O. and Steen, J.B., 1986. Water economy in bird nests. Journal of Comparative

Physiology B 156: 823-828.

Andreev, A.V. and Golubova, E.Y., 1995. A new finding of breeding Kittlitz's Murrelet

Brachyramphus brevirostris on the Ochotsk Sea coast. Russian Journal of 4:

63-64.

Andres, B.A. and Falxa, G.A., 1995. Black (Haematopus bachmani). In: A. Poole

and F. Gill (Editors), The Birds of North America, No. 155. The Academy of Natural

Sciences, Philadelphia, and the American Ornithologists' Union, Washington, D.C., pp.

1-20.

Angielczyk, K.D. and Schmitz, L., 2014. Nocturnality in predates the origin of

by over 100 million . Proceedings of the Royal Society B 281:

10.1098/Rspb.2014.1642.

Antunes, M.T., Taquet, P.V. and Ribeiro, V., 1998. Upper Jurassic dinosaur and crocodile eggs

from Pai Mogo nesting site (Lourinha - Portugal). Memorias da Academia das Ciencias de

Lisboa 37: 83-99.

Ar, A., Paganell, C.V., Reeves, R.B., Greene, D.G. and Rahn, H., 1974. Avian egg: water vapor

conductance, shell thickness, and functional pore area. Condor 76: 153-158.

Ar, A. and Rahn, H., 1978. Interdependence of gas conductance, incubation length, and weight of

the avian egg. In: J. Piiper (Editor), Respiratory Function in Birds, Adult and Embryonic.

Springer-Verlag, Berlin, Germany, pp. 227-235.

Ar, A. and Rahn, H., 1980. Water in the avian egg: overall budget of incubation. American

Zoologist 20: 373-384.

Ar, A. and Rahn, H., 1985. Pores in avian eggshells: gas conductance, gas exchange and 365

embryonic growth rate. Respiration Physiology 61: 1-20.

Ar, A., Rahn, H. and Paganelli, C.V., 1979. Avian egg: mass and strength. Condor 81: 331-337.

Arad, Z., Gavrieli-Levin, I. and Marder, J., 1988. Adaptation of the pigeon egg to incubation in dry

hot environments. Physiological Zoology 61: 293-300.

Arad, Z. and Marder, J., 1981. Egg-shell water vapour conductance of the domestic fowl:

comparison between two breeds and their crosses. British Poultry Science 23: 325-328.

Araujo, R., Castanhinha, R., Martins, R.M.S., Mateus, O., Hendrickx, C., Beckmann, F., Schell, N.

and Alves, L.C., 2013. Filling the gaps of dinosaur eggshell phylogeny:

theropod clutch with embryos from Portugal. Scientific Reports 3: 10.1038/Srep01924.

Archey, G., 1931. Notes on sub-fossil bird remains. Records of the Auckland Institute and

Museum 1: 113-121.

Archey, G., 1941. The moa: a study of the Dinornithiformes. Bulletin of the Auckland Institute and

Museum 1: 1-145.

Argeloo, M. and Dekker, R.W.R.J., 1996. Exploitation of megapode eggs in Indonesia: the role of

traditional methods in the conservation of megapodes. Oryx 30: 59-64.

Asbirk, S., 1979. The adaptive significance of the reproductive pattern in the Black Guillemot,

Cepphus grylle. Videnskaplige Meddelelser Dansk naturhistorisk Forening 141: 29-80.

Baicich, P.J. and Harrison, C.J.O., 1997. A Guide to the Nests, Eggs, and Nestlings of North

American Birds. Academic Press, London, UK, 347 pp.

Baird, P.H., 1994. Black-legged (Rissa tridactyla). In: A. Poole and F. Gill (Editors),

The Birds of North America, No. 92. The Academy of Natural Sciences, Philadelphia, and

the American Ornithologists' Union, Washington, D.C., pp. 1-28.

Baker, A.J. and Coleman, J.D., 1977. The breeding cycle of the Westland Black Petrel (Procellaria

westlandica). Notornis 24: 211-231. 366

Baker, G.C., 1999. Temporal and spatial patterns of laying in the Moluccan megapode Eulipoa

wallacei (G.R. Gray). Zoologische Verhandelingen 327: 53-60.

Baker, G.C. and Dekker, R.W.R.J., 2000. Lunar synchrony in the reproduction of the Moluccan

Megapode Megapodius wallacei. Ibis 142: 382-388.

Balkan, M., Karakas, R. and Biricik, M., 2006. Changes in eggshell thickness, shell conductance

and pore density during incubation in the Peking (Anas platyrhynchos f. dom.). Ornis

Fennica 83: 117-123.

Baltin, S., 1969. Zur Biologie und Ethologie des Talegalla-Huhns (Alectura lathami Gray) unter

besonderer Berücksichtigung des Verhaltens während der Brutperiode. Zeitschrift für

Tierpsychologie 26: 524-572.

Banko, P.C., Black, J.M. and Banko, W.E., 1999. Hawaiian (Nene) (Branta sandvicensis).

In: A. Poole and F. Gill (Editors), The Birds of North America, No. 434. The Academy of

Natural Sciences, Philadelphia, and The American Ornithologists' Union, Washington,

D.C., pp. 1-32.

Barr, J.F., Eberl, C. and McIntyre, J.W., 2000. Red-throated (Gavia stellata). In: A. Poole

and F. Gill (Editors), The Birds of North America, No. 513. The Birds of North America,

Inc., Philadelphia, pp. 1-28.

Barth, E.K., 1953. Calculation of egg volume based on loss of weight during Incubation. Auk 70:

151-159.

Beck, C., 1978. Breeding the West African dwarf crocodile Osteolaemus tetraspis tetraspis at

Memphis Zoo. International Zoo Yearbook 18: 89-91.

Beebe, W., 1931. : Their Lives and Momes, Vols. 1 and 2. Doubleday, Doran and Co.,

New York, 566 pp.

Beetschen, J.-C., Dughi, R. and Sirugue, F., 1977. Sur la presence de coquilles d'œufs d'Oiseaux 367

dans le cretace superieur des Corbieres occidentales. Comptes Rendus de l'Académie des

Sciences-Series II: A-Earth and Planetary Sciences 284: 2491-2494.

Bellrose, F.C., 1976. , Geese, and of North America. Stackpole Books, Harrisburg,

543 pp.

Belopol'skii, L.O., 1961. Ecology of Sea Colony Birds of the Barents Sea. Israel Program for

Scientific Translations, Jerusalem, 346 pp.

Benson, R.B.J., 2010. A description of bucklandii (Dinosauria: Theropoda) from

the Bathonian of the UK and the relationships of Middle Jurassic theropods. Zoological

Journal of the Linnean Society 158: 882-935.

Benson, R.B.J., Campione, N.E., Carrano, M.T., Mannion, P.D., Sullivan, C., Upchurch, P. and

Evans, D.C., 2014. Rates of dinosaur body mass evolution indicate 170 million years of

sustained ecological innovation on the avian stem lineage. PLoS Biology 12:

10.1371/journal.pbio.1001853.

Benson, R.B.J., Carrano, M.T. and Brusatte, S.L., 2010. A new clade of archaic large-bodied

predatory dinosaurs (Theropoda: Allosauroidea) that survived to the latest .

Naturwissenschaften 97: 71-78.

Bertelli, S. and Giannini, N.P., 2005. A phylogeny of extant (Aves: Sphenisciformes)

combining morphology and mitochondrial sequences. Cladistics 21: 209-239.

Bertram, B.C.R. and Burger, A.E., 1981. Aspects of incubation in ostrich. Ostrich 52: 36-43.

Bezuijen, M.R., Cox, J.H., Thorbjarnarson, J.B., Phothitay, C., Hedemark, M. and Rasphone, A.,

2013. Status of Siamese Crocodile (Crocodylus siamensis) Schneider, 1801 (Reptilia:

Crocodylia) in Laos. Journal of Herpetology 47: 41-65.

Birchard, G.F. and Marcellini, D., 1996. Incubation time in reptilian eggs. Journal of Zoology 240:

621-635. 368

Birchard, G.F., Ruta, M. and Deeming, D.C., 2013. Evolution of parental incubation behaviour in

dinosaurs cannot be inferred from clutch mass in birds. Biology Letters 9:

10.1098/Rsbl.2013.0036.

Bland, J.M. and Altman, D.G., 1986. Statistical methods for assessing agreement between two

methods of clinical measurement. Lancet 1: 307-310.

Bland, J.M. and Altman, D.G., 1995. Comparing two methods of clinical measurement: a personal

history. International Journal of Epidemiology 24: S7-S14.

Board, R.G. and Tullett, S.G., 1975. Pore arrangement in Emu (Dromaius novaehollandiae)

eggshell as shown by plastic models. Journal of Microscopy 103: 281-284.

Board, R.G., Tullett, S.G. and Perrott, H.R., 1977. Arbitrary classification of pore systems in avian

eggshells. Journal of Zoology 182: 251-265.

Boersma, P.D., 1982. Why some birds take so long to hatch. American Naturalist 120: 733-750.

Boersma, P.D. and Wheelwright, N.T., 1979. Egg neglect in the Procellariiformes: reproductive

adaptations in the Fork-tailed Storm Petrel. Condor 81: 157-165.

Bolhm, T., 1982. Husbandry of Orinoco crocodiles (Crocodylus intermedius) in Venezuela.

Crocodiles: Proceedings of the 5th Working Meeting of the Crocodile Specialist Group,

IUCN–The World Conservation Union, Gland, Switzerland: 267-285.

Bonde, J.W., Hilton, R.P., Jackson, F.D. and Druschke, P.A., 2015. Fauna of the Newark Canyon

Formation (Lower Cretaceous), east-central Nevada. New Concepts and Discoveries 1-2:

721-732.

Booth, D.T., 1987. Effect of temperature on development of Mallee Fowl Leipoa ocellata eggs.

Physiological Zoology 60: 437-445.

Booth, D.T. and Jones, D.N., 2002. Underground nesting in the megapodes. In: D.C. Deeming

(Editor), Avian Incubation: Behaviour, Environment, and Evolution. Oxford University 369

Press, Oxford, UK, pp. 192-206.

Booth, D.T. and Rahn, H., 1989. Factors modifying rate of water loss from birds eggs during

incubation. Physiological Zoology 63: 697-709.

Booth, D.T. and Seymour, R.S., 1987. Effect of eggshell thinning on water vapor conductance of

Malleefowl eggs. Condor 89: 453-459.

Booth, D.T. and Sotherland, P.R., 1991. Oxygen consumption, air-cell gas tensions and incubation

parameters of Mute eggs. Physiological Zoology 64: 473-484.

Booth, D.T. and Thompson, M.B., 1991. A comparison of reptilian eggs with those of megapode

birds. In: D.C. Deeming and M.W.J. Ferguson (Editors), Egg Incubation: Its Effects on

Embryonic Development in Birds and Reptiles. Cambridge University Press, Cambridge,

UK, pp. 325-344.

Bosque, C. and Bosque, M.T., 1995. Nest predation as a selective factor in the evolution of

developmental rates in altricial birds. American Naturalist 145: 234-260.

Bowen, J., 1996. Notes on the Vanuatu Megapode Megapodius layardi on Ambrym, Vanuatu. Bird

Conservation International 6: 401-408.

Bradley, J.S., Wooller, R.D. and Skira, I.J., 2000. Intermittent breeding in the Short-tailed

Shearwater Puffinus tenuirostris. Journal of Ecology 69: 639-650.

Bravo, A.M. and Gaete, R., 2015. Titanosaur eggshells from the (Upper

Cretaceous, Southern Pyrenees, Spain). Historical Biology 27: 1079-1089.

Bray, E.S., 1999. Eggs and eggshell from the Upper Cretaceous , central

Utah. In: D.D. Gillette (Editor), Vertebrate Paleontology in Utah. Geological Survey

Miscellaneous Publication 99-1, Utah, pp. 361-376.

Bray, E.S. and Lucas, S.G., 1997. Theropod dinosaur eggshell from the Upper Jurassic of New

Mexico. New Mexico Museum of Natural History and Science Bulletin 11: 41-43. 370

Brazaitis, P.J., 1986. Management, reproduction, and growth of Caiman crocodilus yacare at the

New York Zoological Park, Proceedings of the 7th Working Meeting of the Crocodile

Specialist Group of the Species Survival Comission of the International Union for

Conservation of Nature and Natural Resources. IUCN Publication Services, Gland,

Switzerland, pp. 389-397.

Brazaitis, P.J. and Watanabe, M.E., 2011. Crocodilian behaviour: a window to dinosaur behaviour?

Historical Biology 23: 73-90.

Bregulla, H.L., 1992. Birds of Vanuatu. Anthony Nelson, Shropshire, UK, 329 pp.

Brickhill, J., 1986. Breeding success of Malleefowl Leipoa ocellata in central new South Wales.

Emu 87: 42-45.

Broley, C.L., 1947. Migration and nesting of florida bald eagles. Wilson Bulletin 59: 3-20.

Brooke, M., 2004. and Petrels across the World. Oxford University Press, Oxford, UK,

520 pp.

Brown, B. and Schlaikjer, E.M., 1940. The structure and relationships of Protoceratops.

Transactions of the New York Academy of Sciences 2: 99-100.

Brown, C.R., 1994. Nest microclimate, egg temperature, egg water loss, and eggshell conductance

in Cape Weavers Ploceus capensis. Ostrich 65: 26-31.

Brown, C.R. and Adams, N.J., 1988. Egg temperature, embryonic metabolism, and water loss

from the eggs of subantarctic Procellariiformes. Physiological Zoology 61: 126-136.

Brua, R.B., 2001. Ruddy Duck (Oxyura jamaicensis). In: A. Poole and F. Gill (Editors), The Birds

of North America, No. 696. The Birds of North America, Inc., Philadelphia, pp. 1-32.

Buckley, P.A. and Buckley, F.G., 2002. Royal (Sterna maxima). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 700. The Birds of North America, Inc.,

Philadelphia, pp. 1-28. 371

Bunce, M., Worthy, T.H., Phillips, M.J., Holdaway, R.N., Willerslev, E., Haile, J., Shapiro, B.,

Scofield, R.P., Drummond, A., Kamp, P.J.J. and Cooper, A., 2009. The evolutionary

history of the extinct ratite moa and New Zealand Neogene paleogeography. Proceedings

of the National Academy of Sciences of the United States of America 106: 20646-20651.

Burger, J. and Gochfeld, M., 1987. Nest-site selection by Mew (Larus canus): a comparison

of marsh and dry-land colonies. Wilson Bulletin 99: 673-687.

Burke, E., 1925. A study of incubation. Bulletin of University of Montana Agricultural

Experimental Station 178: 1-43.

Burness, G.P., Lefevre, K. and Collins, C.T., 1999. Elegant Tern (Sterna elegans). In: A. Poole and

F. Gill (Editors), The Birds of North America, No. 404. The Birds of North America, Inc.,

Philadelphia, pp. 1-28.

Burton, F.G. and Tullett, S.G., 1983. A comparison of the effects of eggshell porosity on the

respiration and growth of domestic fowl, duck and turkey embryos. Comparative

Biochemistry and Physiology A 75: 167-174.

Bustard, H.R., 1980. Clutch size, incubation and hatching success of [Gavialis gangeticus

(Gmelin)] eggs from Narayani River, Nepal, 1976-1978. Journal of Bombay Natural

History Society 77: 100-105.

Butler, M.A. and King, A.A., 2004. Phylogenetic comparative analysis: a modeling approach for

adaptive evolution. American Naturalist 164: 683-695.

Butler, R.W., Verbeek, N.A.M. and Richardson, H., 1984. The breeding biology of the

Northwestern Crow. Wilson Bulletin 96: 408-418.

Buttemer, W.A., Astheimer, L.B. and Dawson, T.J., 1988. Thermal and water relations of emu eggs

during natural incubation. Physiological Zoology 61: 483-494.

Calder III, W.A., 1978. The Kiwi: a case of compensating divergences from allometric predictions. 372

In: J. Piiper (Editor), Respiratory Function in Birds, Adult and Embryonic. Springer-Verlag,

Berlin, Germany, pp. 239-242.

Calder III, W.A., 1979. Kiwi and egg design: evolution as a package deal. BioScience 29: 461-467.

Calder III, W.A., Parr, C.R. and Karl, D.P., 1978. Energy content of eggs of Brown Kiwi Apteryx

australis; extreme in avian evolution. Comparative Biochemistry and Physiology A 60:

177-179.

Calvo, J.O., Engelland, S., Heredia, S.E. and Salgado, L., 1997. First record of dinosaur eggshells

(? - Megaloolithidae) from Neuquen, Patagonia, Argentina. Gaia 14: 23-32.

Campbell, H.W., 1972. Ecological or phylogenetic interpretations of crocodilian nesting habits.

Nature 238: 404-405.

Campione, N.E. and Evans, D.C., 2012. A universal scaling relationship between body mass and

proximal limb bone dimensions in quadrupedal terrestrial tetrapods. BMC Biology 10:

10.1186/1741-7007-10-60.

Canudo, J.I., Gasca, J.M., Aurell, M., Badiola, A., Blain, H.A., Cruzado-Caballero, P.,

Gomez-Fernandez, D., Moreno-Azanza, M., Parrilla, J., Rabal-Garces, R. and

Ruiz-Omenaca, J.I., 2010. La Cantalera: an exceptional window onto the vertebrate

biodiversity of the - transition in the Iberian Peninsula. Journal of

Iberian Geology 36: 205-224.

Carey, C., 1979. Increase in conductance to water vapor during incubation in eggs of two avian

species. Journal of Experimental Zoology 209: 181-186.

Carey, C., 1980. The ecology of avian incubation. Bioscience 30: 819-824.

Carey, C., 1983. Structure and function of avian eggs. In: R.F. Johnston (Editor), Current

Ornithology. Plenum Press, New York, pp. 69-103.

Carey, C., 2002. Incubation in extreme environments. In: D.C. Deeming (Editor), Avian 373

Incubation: Behaviour, Environment, and Evolution. Oxford University Press, Oxford, UK,

pp. 238-253.

Carey, C., Garber, S.D., Thompson, E.L. and James, F.C., 1983. Avian reproduction over an

altitudinal gradient. II. Physical characteristics and water loss of eggs. Physiological

Zoology 56: 340-352.

Carey, C., Leon-Velarde, F., Dunin-Borkowski, O., Bucher, T.L., de La Torre, G., Espinoza, D. and

Monge, C., 1989. Variation in eggshell characteristics and gas exchange of montane and

lowland coot eggs. Journal of Comparative Physiology B 159: 389-400.

Carey, C., Leon-Velarde, F., Dunin-Borkowskit, O. and Monge, C., 1989. Shell conductance, daily

water loss, and water content of Puna Teal eggs. Physiological Zoology 62: 83-95.

Carey, C., Leonvelarde, F., Castro, G. and Monge, C., 1987. Shell conductance, daily water loss,

and water content of Andean Gull and Puna Ibis eggs. Journal of Experimental Zoology:

247-252.

Carey, C., Leonvelarde, F. and Monge, C., 1990. Eggshell conductance and other physical

characteristics of avian eggs laid in the Peruvian . Condor 92: 790-793.

Carpenter, K., 1999. Eggs, Nests, and Baby Dinosaurs: a Look at Dinosaur Reproduction. Indiana

University Press, Bloomington, 336 pp.

Carrano, M., Mateus, O. and Mitchell, J., 2013. First definitive association between embryonic

Allosaurus bones and Prismatoolithus eggs in the Morrison Formation (Upper Jurassic,

Wyoming, USA). Journal of Vertabrate Paleontology 33. Conference abstracts: 101.

Carrano, M.T., Benson, R.B.J. and Sampson, S.D., 2012. The phylogeny of (Dinosauria:

Theropoda). Journal of Systematic Palaeontology 10: 211-300.

Castanet, J., Rogers, K.C., Cubo, J. and Boisard, J.J., 2000. Periosteal bone growth rates in extant

ratites (ostriche and emu). Implications for assessing growth in dinosaurs. Comptes 374

Rendus De L Academie Des Sciences Serie III-Sciences De La Vie-Life Sciences 323:

543-550.

Cavitt, J.F. and Haas, C.A., 2000. (Toxostoma rufum). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 557. The Birds of North America, Inc.,

Philadelphia, pp. 1-28.

Chabreck, R.H., 1973. Temperature variation in nests of the American Alligator. Herpetologica 29:

48-51.

Chabreck, R.H., 1975. Moisture variation in nests of the American Alligator (Alligator

mississippiensis). Herpetologica 31: 385-389.

Charruau, P., 2012. Microclimate of American Crocodile nests in Banco Chinchorro biosphere

reserve, Mexico: effect on incubation length, embryos survival and hatchlings sex. Journal

of Thermal Biology 37: 6-14.

Charruau, P., Thorbjarnarson, J.B. and Henaut, Y., 2010. Tropical cyclones and reproductive

ecology of Crocodylus acutus Cuvier, 1807 (Reptilia: : Crocodylidae) on a

Caribbean atoll in Mexico. Journal of Natural History 44: 741-761.

Chassagne-Manoukian, M., Haddoumi, H., Cappetta, H., Charriere, A., Feist, M., Tabuce, R. and

Vianey-Liaud, M., 2013. Dating the "red beds" of the Eastern Moroccan High Plateaus:

evidence from late Late Cretaceous charophytes and dinosaur eggshells. Geobios 46:

371-379.

Chattock, A.P., 1925. On the physics of incubation. Philosophical Transactions of the Royal

Society of London Series B: Containing Papers of a Biological Character 213: 397-450.

Cheng, Y.N., Jui, Q., Wu, X.C. and Shou, H.Y., 2008. Oviraptorosaurian eggs (Dinosauria) with

embryonic skeletons discovered for the first time in China. Acta Geologica Sinica 82:

1089-1094. 375

Chiappe, L.M., Coria, R.A., Dingus, L., Jackson, F., Chinsamy, A. and Fox, M., 1998. Sauropod

dinosaur embryos from the late Cretaceous of Patagonia. Nature 396: 258-261.

Chiappe, L.M., Dingus, L., Jackson, F.D., Grellet-Tinner, G., Coria, R., Loope, D., Clarke, J. and

Garrido, A., 2001. Sauropod eggs and embryos from the Late Cretaceous of Patagonia. In:

A.M. Bravo and T. Reyes (Editors), First International Symposium on Dinosaur Eggs and

Babies Extended Abstracts, Catalonia, Spain, pp. 23-29.

Chiappe, L.M., Schmitt, J.G., Jackson, F.D., Garrido, A., Dingus, L. and Grellet-Tinner, G., 2004.

Nest structure for sauropods: sedimentary criteria for recognition of dinosaur nesting traces.

Palaios 19: 89-95.

Choiniere, J.N., Xu, X., Clark, J.M., Forster, C.A., Guo, Y. and Han, F.L., 2010. A basal

alvarezsauroid theropod from the Early Late Jurassic of Xinjiang, China. Science 327:

571-574.

Chow, M.M., 1951. Notes on the Late Cretaceous dinosaurian remains and the fossil eggs from

Laiyang Shantung. Bulletin of the Geological Society of China 31: 89-96.

Chow, M.M., 1954. On the microstructure of the fossil egg-shells from Laiyang, Shantung. Acta

Palaeontologica Sinica 2: 389-394.

Cintra, R., 1988. Nesting ecology of the Paraguayan Caiman (Caiman yacare) in the Brazilian

Pantanal. Journal of Herpetology 22: 219-222.

Clark, J.M., Norell, M.A. and Chiappe, L.M., 1999. An oviraptorid skeleton from the Late

Cretaceous of Ukhaa Tolgod, Mongolia, preserved in an avianlike brooding position over

an oviraptorid nest. American Museum Novitates 3265: 1-35.

Clark, M.E., Boonstra, T.A., Reed, W.L. and Gastecki, M.L., 2010. Intraclutch variation in egg

conductance facilitates hatching synchrony of Canada Geese. Condor 112: 447-454.

Clayburn, J.K., Smith, D.L. and Hayward, J.L., 2004. Taphonomic effects of pH and temperature 376

on extant avian dinosaur eggshell. Palaios 19: 170-177.

Close, R.A. and Rayfield, E.J., 2012. Functional morphometric analysis of the furcula in mesozoic

birds. PLoS One 7: 10.1371/journal.pone.0036664.

Codrea, V., Smith, T., Dica, P., Folie, A., Garcia, G., Godefroit, P. and Van Itterbeeck, J., 2002.

Dinosaur egg nests, mammals and other vertebrates from a new Maastrichtian site of the

Hateg Basin (Romania). Comptes Rendus Palevol 1: 173-180.

Colacino, J.M., Hamel, P.B., Rahn, H., Board, R.G. and Sparks, N.H.C., 1985. The vaterite cover

on the eggs of Anhinga anhinga and its effect on gas conductance (Aves: Anhingidae).

Journal of Zoology 205: 425-433.

Colbourne, R., 2002. Incubation behaviour and egg physiology of Kiwi (Apteryx spp.) in natural

habitats. New Zealand Journal of Ecology 26: 129-138.

Coombs, W.P.J., 1989. Modern analogues for dinosaur nesting and parenting behavior. In: J.O.

Farlow (Editor), Paleobiology of the Dinosaurs. Geological Society of America, Boulder,

pp. 21-53.

Cornejo, J., Iorizzo, M. and Clum, N., 2014. Artificial incubation of Maleo Macrocephalon maleo

eggs at the Bronx Zoo/Wildlife Conservation Society, New York. International Zoo

Yearbook 48: 39-47.

Cott, H.B., 1961. Scientific results of an inquiry into the ecology and economic status of the Nile

Crocodile (Crocodylus niloticus) in Uganda and Northern Rhodesia. The Transactions of

the Zoological Society of London 29: 211-356.

Cousin, R., Breton, G., Fournier, R. and Watte, J.-P., 1989. Dinosaur egg-laying and nesting: the

case of an Upper Maastrichtian site at Rennes-le-Chateau (Aude, France). Historical

Biology 2: 157-167.

Cramp, S., 1977. Handbook of the Birds of Europe, the Middle East, and North Africa: the Birds of 377

the Western Palearctic. Vol. 1: Ostrich to Ducks. Oxford University Press, Oxford, UK,

722 pp.

Cramp, S., 1980. Handbook of the Birds of Europe, the Middle East, and North Africa: the Birds of

the Western Palearctic. Vol. 2: Hawks to . Oxford University Press, Oxford, UK,

695 pp.

Cramp, S., 1983. Handbook of the Birds of Europe, the Middle East, and North Africa: the Birds of

the Western Palearctic. Vol. 3: Warders to Gulls. Oxford University Press, Oxford, UK,

1000 pp.

Cramp, S., 1985. Handbook of the Birds of Europe, the Middle East, and North Africa: the Birds of

the Western Palearctic. Vol. 4: to . Oxford University Press, Oxford,

UK, 960 pp.

Cramp, S., 1988. Handbook of the Birds of Europe, the Middle East, and North Africa: the Birds of

the Western Palearctic. Vol. 5: Tyrant Flycatchers to Thrushes. Oxford University Press,

Oxford, UK, 1136 pp.

Crawshaw, P.G. and Schaller, G.B., 1980. Nesting of Paraguayan Caiman (Caiman yacare) in

Brazil. Papeis Avulsos de Zoologia 33: 283-292.

Crome, F.H.J. and Brown, H.E., 1979. Notes on social organization and breeding of the

Orange-footed Megapodius reinwardt. Emu 79: 111-119.

Crowe, T.M., Bowie, R.C.K., Bloomer, P., Mandiwana, T.G., Hedderson, T.A.J., Randi, E., Pereira,

S.L. and Wakeling, J., 2006. Phylogenetics, biogeography and classification of, and

character evolution in, gamebirds (Aves: Galliformes): effects of character exclusion, data

partitioning and missing data. Cladistics 22: 495-532.

Cunha, P., Mateus, P.O. and Antunes, M.T., 2004. The sedimentology of the Paimogo dinosaur

nest site (Portugal, Upper Jurassic). 23rd IAS Meeting of Sedimentology Abstract Book: 378

93.

Davies, S.J.J.F., 2002. Ratites and Tinamous: Tinamidae, , Dromaiidae, Casuariidae,

Apterygidae, . Oxford University Press, Oxford, UK, 310 pp.

Davis, J., W.E., 1993. Black-crowned Night-Heron (Nycticorax nycticorax). In: A. Poole and F.

Gill (Editors), The Birds of North America, No. 74. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp. 1-20.

Davis, T.A. and Ackerman, R.A., 1985. Adaptations of Black Tern (Chilidonias niger) eggs for

water loss in a moist nest. Auk 102: 640-643.

Davis, T.A., Platter-Reiger, M.F. and Ackerman, R.A., 1984. Incubation water loss by Pied-billed

Grebe: eggs adaptation to a hot wet nest. Physiological Zoology 57: 384-391.

Davis, W.E. and Kricher, J., 2000. Glossy Ibis (Plegadis falcinellus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 545. The Birds of North America, Inc.,

Philadelphia, pp. 1-20.

Dawson, D.G., 1964. The eggs of the House Sparrow. Notornis 11: 187-189. de Kloet, R.S. and de Kloet, S.R., 2005. The evolution of the spindlin gene in birds: sequence

analysis of an intron of the spindlin W and Z gene reveals four major divisions of the

Psittaciformes. Molecular Phylogenetics and Evolution 36: 706-721. de Ricqlès, A., Mateus, O., Antunes, M.T. and Taquet, P., 2001. Histomorphogenesis of embryos

of Upper Jurassic theropods from Lourinhã (Portugal). Comptes Rendus de l'Académie des

Sciences-Series II: A-Earth and Planetary Sciences 332: 647-656.

Deeming, D.C., 2002a. Behaviour patterns during incubation. In: D.C. Deeming (Editor), Avian

Incubation: Behavior, Environment, and Evolution. Oxford University Press, Oxford, UK,

pp. 63-87.

Deeming, D.C., 2002b. Functional characteristics of eggs. In: D.C. Deeming (Editor), Avian 379

Incubation: Behaviour, Environment, and Evolution. Oxford University Press, Oxford, UK,

pp. 28-42.

Deeming, D.C., 2002c. Importance and evolution of incubation in avian reproduction. In: D.C.

Deeming (Editor), Avian Incubation: Behaviour, Environment, and Evolution. Oxford

University Press, Oxford, UK, pp. 1-7.

Deeming, D.C., 2006. Ultrastructural and functional morphology of eggshells supports the idea

that dinosaur eggs were incubated buried in a substrate. Palaeontology 49: 171-185.

Deeming, D.C., Birchard, G.F., Crafer, R. and Eady, P.E., 2006. Egg mass and incubation period

allometry in birds and reptiles: effects of phylogeny. Journal of Zoology 270: 209-218.

Deeming, D.C. and Ferguson, M.W.J., 1989. Effects of incubation temperature on growth and

development of embryos of Alligator mississippiensis. Journal of Comparative Physiology

B-Biochemical Systemic and Environmental Physiology 159: 183-193.

Deeming, D.C. and Ruta, M., 2014. Egg shape changes at the theropod-bird transition, and a

morphometric study of eggs. Royal Society Open Science 1:

10.1098/Rsos.140311.

Dekker, R.W.R.J., 1988. Notes on ground temperatures at nesting sites of the Maleo

Macrocephalon maleo (Megapodiidae). Emu 88: 124-127.

Dekker, R.W.R.J., Argeloo, M. and Jepson, P., 1995. Notes on the Moluccan Megapode Eulipoa

wallacei (G.R. Gray, 1860) following the rediscovery of two major nesting grounds.

Zoologische Mededelingen 69: 251-260.

Dekker, R.W.R.J. and Brom, T.G., 1992. Megapode phylogeny and the interpretation of incubation

strategies. Zoologische Verhandelingen 278: 19-31. del Hoyo, J., Elliott, A. and Sargatal, J., 1992. Handbook of the Birds of the World, Vol. 1: Ostrich

to Ducks. Lynx Edicions, Barcelona, Spain, 696 pp. 380

del Hoyo, J., Elliott, A. and Sargatal, J., 1994. Handbook of the Birds of the World, Vol. 2: New

World Vultures to . Lynx Edicions, Barcelona, Spain, 638 pp. del Hoyo, J., Sargatal, J. and Christie, D.A., 2011. Handbook of the Birds of the World, Vol. 16:

Tanagers to New World Blackbirds. Lynx Edicions, Barcelona, Spain, 800 pp.

Delacour, J., 1959. The Waterfowl of the World, Vol. 3. McMlix, London, UK, 270 pp.

Dewaele, L., Tsogtbaatar, K., Barsbold, R., Garcia, G., Stein, K., Escuillie, F. and Godefroit, P.,

2015. Perinatal Specimens of Saurolophus angustirostris (Dinosauria: Hadrosauridae),

from the Upper Cretaceous of Mongolia. PLoS One 10: 10.1371/journal.pone.0138806.

Diaz-Molina, M., Kalin, O., Benito, M.I., Lopez-Martinez, N. and Vicens, E., 2007. Depositional

setting and early diagenesis of the dinosaur eggshell-bearing Aren Fm at Bastus, Late

Campanian, south-central Pyrenees. Sedimentary Geology 199: 205-221.

Donaire, M. and Lopez-Martinez, N., 2009. Porosity of Late Ornitholithus eggshells

(Tremp Fm, south-central Pyrenees, Spain): palaeoclimatic implications.

Palaeoclimatology Palaeoecology 279: 147-159.

Dong, Z.M. and Currie, P.J., 1996. On the discovery of an oviraptorid skeleton on a nest of eggs at

Bayan Mandahu, Inner Mongolia, People's Republic of China. Canadian Journal of Earth

Sciences 33: 631-636.

Drent, R., 1975. Incubation. In: D.S. Farner and J.R. King (Editors), Avian Biology. Academic

Press, New York, pp. 333-420.

Drent, R.H., 1970. Functional aspects of incubation in the Herring Gull. Behaviour. Supplement

17: 1-132.

Drilling, N., Titman, R. and McKinney, F., 2002. (Anas platyrhynchos). In: A. Poole and F.

Gill (Editors), The Birds of North America, No. 658. The Birds of North America, Inc.,

Philadelphia, pp. 1-44. 381

Dugger, B.D., Dugger, K.M. and Fredrickson, L.H., 1994. Hooded Merganser (Lophodytes

cucullatus). In: A. Poole and F. Gill (Editors), The Birds of North America, No. 98. The

Academy of Natural Sciences, Philadelphia, and the American Ornithologists' Union,

Washington, D.C., pp. 1-24.

Dunn, E.H. and Agro, D.J., 1995. Black Tern (Chlidonias niger). In: A. Poole and F. Gill (Editors),

The Birds of North America, No. 147. The Academy of Natural Sciences, Philadelphia,

and The American Ornithologists' Unnion, Washington, D.C., pp. 1-24.

Eadie, J.M., Mallory, M.L. and Lumsden, H.G., 1995. Common Goldeneye (Bucephala clangula).

In: A. Poole and F. Gill (Editors), The Birds of North America, No. 170. The Academy of

Natural Sciences, Philadelphia, and the American Ornithologists' Union, Washington, D.C.,

pp. 1-32.

Eadie, J.M., Savard, J.-P.L. and Mallory, M.L., 2000. Barrow's Goldeneye (Bucephala islandica).

In: A. Poole and F. Gill (Editors), The Birds of North America, No. 548. The Birds of North

America, Inc., Philadelphia, pp. 1-32.

Eagle, R.A., Enriquez, M., Grellet-Tinner, G., Perez-Huerta, A., Hu, D., Tutken, T., Montanari, S.,

Loyd, S.J., Ramirez, P., Tripati, A.K., Kohn, M.J., Cerling, T.E., Chiappe, L.M. and Eiler,

J.M., 2015. Isotopic ordering in eggshells reflects body temperatures and suggests

differing thermophysiology in two Cretaceous dinosaurs. Nature Communications 6:

10.1038/ncomms9296.

Eaton, S.W., 1992. Wild Turkey. In: A. Poole, P. Stettenheim and F. Gill (Editors), The Birds of

North America, No. 22. The Academy of Natural Sciences, Philadelphia, and The

American Ornithologists' Union, Washington, D.C., pp. 1-28.

Elliot, R.D. and Morrison, R.I.G., 1979. Incubation period of the Yellow . Auk 96: 422-423.

Ely, C.R. and Dzubin, A.X., 1994. Greater White-fronted Goose (Anser albifrons). In: A. Poole 382

and F. Gill (Editors), The Birds of North America, No. 131. The Academy of Natural

Sciences, Philadelphia, and the American Ornithologists' Union, Washington, D.C., pp.

1-32.

Ely, C.R. and Raveling, D.G., 1984. Breeding biology of Pacific White-fronted Geese. Journal of

Wildlife Management 48: 823-837.

Ensom, P.C., 1996. eggshell from the Purbeck Group of Dorset, Southern

England. Proceedings of Dorset Natural History and Archaeological Society 118: 79-83.

Erben, H.K., 1970. Ultrastrukturen und Mineralisation rezenter und fossiler Eischalen bei Vogeln

und Reptilien. Biomineralisation Forschsber 6: 32-48.

Erben, H.K., Hoefs, J. and Wedepohl, K.H., 1979. Paleobiological and isotopic studies of

eggshells from a declining dinosaur species. Paleobiology 5: 380-414.

Ericson, P.G.P., Anderson, C.L., Britton, T., Elzanowski, A., Johansson, U.S., Kallersjo, M.,

Ohlson, J.I., Parsons, T.J., Zuccon, D. and Mayr, G., 2006. Diversification of :

integration of molecular sequence data and fossils. Biology Letters 2: 543-547.

Espinosa-Blanco, A.S., Seijas, A.E. and Hernández, O., 2013. Egg collection as a conservation

tool of Orinoco Crocodile (Crocodylus intermedius) in the Cojedes River System,

Venezuela. Interciencia 38: 358.

Ewins, P.J., 1993. Pigeon Gullemot (Cepphus columba). In: A. Poole and F. Gill (Editors), The

Birds of North America, No. 49. The Academy of Natural Sciences, Philadelphia, and The

American Ornithologists' Union, Washington, D.C., pp. 1-24.

Faccio, G., 1994. Dinosaurian eggs from the Upper Cretaceous of Uruguay. In: K. Carpenter, K.F.

Hirsch and J.R. Horner (Editors), Dinosaur Eggs and Babies. Cambridge University Press,

Cambridge, UK, pp. 47-55.

Falanruw, M.V.C., 1975. Distribution of the Micronesian Megapode Megapodius laperouse in the 383

Northern Mariana Islands. Micronesica 11: 149-150.

Fang, X., Lu, L., Cheng, Z., Zou, Y., Pang, Q., Wang, Y., Chen, K., Yin, Z., Wang, X., Liu, J., Xie,

H. and Jin, Y., 1998. On the Cretaceous Fossil Eggs of Xixia County, Henan Province.

Geological Publishing House, Beijing, China, 125 pp.

Fang, X., Lu, L., Jiang, Y. and Yang, L., 2003. Cretaceous fossil eggs from the Tiantai basin of

Zhejiang, with a discussion on the extinction of dinosaurs. Geological Bulletin of China

22: 512-520.

Fang, X., Wang, Y. and Jiang, Y., 2000. On the Late Cretaceous fossil eggs of Tiantai, Zhejiang.

Geological Review 46: 107-112.

Fang, X., Zhang, Z., Pang, Q., Li, P.X., Han, Y., Xie, H., Yan, R., Pang, F., Lu, J. and Cheng, Z.,

2007. Cretaceous strata and egg fossils in Xixia, Henan Province. Acta Geologica Sinica

28: 123-142.

Fang, X., Zhang, Z., Zhang, X., Lu, L., Han, Y. and Li, P., 2005. Fossil eggs from the Heyuan

Basin, east-central Guangdong, China. Geological Bulletin of China 24: 682-686.

Fang, X.S., Li, P.X., Zhang, Z.J., Zhang, X.Q., Lin, Y.L., Guo, S.B., Cheng, Y.M., Li, Z.Y., Zhang,

X.J. and Cheng, Z.W., 2009. Cretaceous strata in Nanxiong Basin of Guangdong and the

evolution from the dinosaur egg to the . Acta Geoscientica Sinica 30: 167-186.

Fanti, F., Currie, P.J. and Badamgarav, D., 2012. New specimens of Nemegtomaia from the

Baruungoyot and Nemegt formations (Late Cretaceous) of Mongolia. PLoS One 7:

10.1371/journal.pone.0031330.

Felsenstein, J., 1985. Phylogenies and the comparative method. American Naturalist 125: 1-15.

Felsenstein, J., 1988. Phylogenies and quantitative characters. Annual Review of Ecology and

Systematics 19: 445-471.

Ferguson, M.W.J., 1985. Reproductive biology and embryology of the crocodilians. In: C. Gans 384

(Editor), Biology of the Reptilia. John Wiley and Sons, New York, pp. 329-491.

Fernandez, M.S. and Khosla, A., 2015. Parataxonomic review of the Upper Cretaceous dinosaur

eggshells belonging to the oofamily Megaloolithidae from India and Argentina. Historical

Biology 27: 158-180.

Fiorelli, L.E., Grellet-Tinner, G., Alasino, P.H. and Arganaraz, E., 2012. The geology and

palaeoecology of the newly discovered Cretaceous neosauropod hydrothermal nesting site

in Sanagasta (Los Llanos Formation), La Rioja, northwest Argentina. Cretaceous Research

35: 94-117.

Fiorillo, A.R., Hasiotis, S.T. and Kobayashi, Y., 2014. Herd structure in Late Cretaceous polar

dinosaurs: a remarkable new dinosaur tracksite, Denali National Park, Alaska, USA.

Geology 42: 719-722.

Fisher, H.I., 1969. Eggs and egg-laying in the Laysan , Diomedea immutabilis. Condor

71: 102-112.

Fitzherbert, K., 1985. The role of energetic factors in the evolution of the breeding biology of the

Short-tailled Shearwater (Puffinus tenuirostris, Temminck). Ph.D. Thesis, Monash

University, Victoria, Australia, 225 pp.

Fjeldsa, J., 2004. The Grebes: Podicipedidae. Oxford University Press, Oxford, UK, 264 pp.

Fleay, D.H., 1937. Nesting habits of the Brush Turkey. Emu 36: 153-163.

Fondevilla, V., Dinares-Turell, J., Vila, B., Le Loeuff, J., Estrada, R., Oms, O. and Galobart, A.,

2016. Magnetostratigraphy of the Maastrichtian continental record in the Upper Aude

Valley (northern Pyrenees, France): Placing age constraints on the succession of

dinosaur-bearing sites. Cretaceous Research 57: 457-472.

Foster, N.H., 1902. Observations on the weights of birds' eggs. The Irish Naturalist 1902: 237-245.

Frederick, P.C., 1997. Tricolored Heron (Egretta tricolor). In: A. Poole and F. Gill (Editors), The 385

Birds of North America, No. 306. The Academy of Natural Sciences, Philadelphia, and

The American Ornithologists' Union, Washington, D.C., pp. 1-28.

Frederick, P.C. and Siegel-Causey, D., 2000. Anhinga (Anhinga anhinga). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 522. The Birds of North America, Inc.,

Philadelphia, pp. 1-24.

French, N.A. and Board, R.G., 1983. Water vapour conductance of wildfowl eggs and incubator

humidity. Wildfowl 34: 144-152.

Friedlander, B., 1899. Über die Nestlöcher des Megapodius pritchardi auf der Insel Niuafu.

Ornithologische Monatsberichte 7: 37-40.

Frith, C.B. and Frith, D.W., 1995. Notes on the nesting biology and diet of Victoria's Riflebird

Ptiloris victoriae. Emu 95: 162-174.

Frith, C.B. and Frith, D.W., 1998. Nesting biology of the Golden Bowerbird Prionodura

newtoniana endemic to Australian upland tropical rainforest. Emu 98: 245-268.

Frith, D.W. and Frith, C.B., 2000. The nesting biology of the Grey-headed Robin Heteromyias

albispecularis (Petroicidae) in Australian upland tropical rainforest. Emu 100: 81-94.

Frith, H.J., 1956. Breeding habits in the family Megapodiidae. Ibis 98: 620-640.

Frith, H.J., 1959. Breeding of the Mallee Fowl, Leipoa ocellata Gould (Megapodiidae). Wildlife

Research 4: 31-60.

Fritz, S.A. and Purvis, A., 2010. Selectivity in mammalian extinction risk and threat types: a new

measure of phylogenetic signal strength in binary traits. Conservation Biology 24:

1042-1051.

Garcia, G., 2000. New data on the fossil eggshell diversity from the lower Maastrichtian of

southern France. Geobios 33: 113-126.

Garcia, G., Feist, M., Cabot, A., Valentin, X. and Vianey-Liaud, M., 2000. New Prismatoolithus 386

oospecies and other Upper Cretaceous dinosaur eggs from the Villeveyrac-Meze Basin

(Herault, France). Bulletin de la Societe Geologique de France 171: 283-289.

Garcia, G., Tabuce, R., Cappetta, H., Marandat, B., Bentaleb, I., Benabdallah, A. and

Vianey-Liaud, M., 2003. First record of dinosaur eggshells and teeth from the North-West

African Maastrichtian (Morocco). Palaeovertebrata 32: 59-69.

Garcia, G. and Vianey-Liaud, M., 2001. Dinosaur eggshells as biochronological markers in Upper

Cretaceous continental deposits. Palaeogeography Palaeoclimatology Palaeoecology 169:

153-164.

Garcia, R.A. and Cerda, I.A., 2010. Dentition and histology in titanosaurian dinosaur embryos

from Upper Cretaceous of Patagonia, Argentina. Palaeontology 53: 335-346.

Garland, T., Bennett, A.F. and Rezende, E.L., 2005. Phylogenetic approaches in comparative

physiology. Journal of Experimental Biology 208: 3015-3035.

Garland, T., Harvey, P.H. and Ives, A.R., 1992. Procedures for the analysis of comparative data

using phylogenetically independent contrasts. Systematic Biology 41: 18-32.

Garland, T. and Ives, A.R., 2000. Using the past to predict the present: confidence intervals for

regression equations in phylogenetic comparative methods. American Naturalist 155:

346-364.

Gartner, G.E.A., Hicks, J.W., Manzani, P.R., Andrade, D.V., Abe, A.S., Wang, T., Secor, S.M. and

Garland, T., 2010. Phylogeny, ecology, and heart position in . Physiological and

Biochemical Zoology 83: 43-54.

Gaston, A.J., 1994. Ancient Murrelet (Synthliboramphus antiquus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 132. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Unnion, Washington, D.C., pp. 1-24.

Gaston, A.J. and Dechesne, S.B.C., 1996. Rhinoceros Auklet (Cerorhinca monocerata). In: A. 387

Poole and F. Gill (Editors), The Birds of North America, No. 212. The Academy of Natural

Sciences, Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp.

1-20.

Gaston, A.J. and Jones, I.L., 1998. The Auks. Oxford University Press, Oxford, UK, 388 pp.

Gaston, A.J. and Nettleship, D.N., 1981. The thick-billed Murres of Prince Leopold Island: a study

of the breeding ecology of a colonial high arctic seabirds. Monograph Series No. 6.

Canadian Wildlife Service, Ottawa, Canada, 350 pp.

Geist, N.R. and Jones, T.D., 1996. Juvenile skeletal structure and the reproductive habits of

dinosaurs. Science 272: 712-714.

Genise, J.F. and Sarzetti, L.C., 2011. Fossil cocoons associated with a dinosaur egg from Patagonia,

Argentina. Palaeontology 54: 815-823.

Gibb, G.C., Kennedy, M. and Penny, D., 2013. Beyond phylogeny: pelecaniform and ciconiiform

birds, and long-term niche stability. Molecular Phylogenetics and Evolution 68: 229-238.

Gill, B.J., 2000. Morphometrics of moa eggshell fragments (Aves: Dinornithiformes) from Late

Holocene -sands of the Karikari Peninsula, New Zealand. Journal of the Royal

Society of New Zealand 30: 131-145.

Gill, B.J., 2006. A catalogue of moa eggs (Aves: Dinornithiformes). Records of the Auckland

Museum 43: 55-80.

Gill, F.B., 2007. Ornithology. W.H. Freeman and Company, New York, 758 pp.

Gilliard, E.T. and LeCroy, M., 1970. Notes on birds from the Tamrau Mountains, New Guinea.

American Museum Novitates 2420: 1-28.

Giudice, J.H. and Ratti, J.T., 2001. Ring-necked (Phasianus colchicus). In: A. Poole and

F. Gill (Editors), The Birds of North America, No. 572. The Birds of North America, Inc.,

Philadelphia, pp. 1-32. 388

Gladstone, H.S., 1904. Note on the decrease in the weight of eggs as incubation advances. Ibis 28:

376.

Glass, P.O., Reichel, J.D., Villagomez, E.C., Taisacan, E. and Aldan, D., 1988. Micronesian

megapode surveys and research. Five Year Progress Report 1983-1987

(Pittman-Robertson Federal Aid in Wildlife Restoration Program), pp. 131-153.

Gochfeld, M. and Burger, J., 1994. Black Skimmer (Rynchops niger). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 108. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp. 1-28.

Godefroit, P., Golovneva, L., Shchepetov, S., Garcia, G. and Alekseev, P., 2009. The last polar

dinosaurs: high diversity of latest Cretaceous arctic dinosaurs in Russia.

Naturwissenschaften 96: 495-501.

Golovneva, L., 2000. The Maastrichtian (Late Cretaceous) climate in the Northern Hemisphere.

Geological Society, London, Special Publications 181: 43-54.

Gonzalez, J., Duttmann, H. and Wink, M., 2009. Phylogenetic relationships based on two

mitochondrial genes and hybridization patterns in . Journal of Zoology 279:

310-318.

Good, T.P., 1998. Great Black-backed Gull (Larus marinus). In: A. Poole and F. Gill (Editors), The

Birds of North America, No. 330. The Birds of North America, Inc., Philadelphia, pp. 1-32.

Goodfellow, P., 1977. Birds as Builders. David and Charles, London, UK, 168 pp.

Goodwin, T.M. and Marion, W.R., 1978. Aspects of nesting ecology of American

(Alligator mississippiensis) in north-central Florida. Herpetologica 34: 43-47.

Goth, A. and Vogel, U., 1996-1997. Egg laying and incubation of the Polynesian Megapodes.

Annual Review of the World Pheasant Association: 43-54.

Gottfried, M.D., O'Connor, P.M., Jackson, R.D., Roberts, E.M. and Cham, R., 2004. Dinosaur 389

eggshell from the red sandstone group of Tanzania. Journal of Vertebrate Paleontology 24:

494-497.

Goudie, R.I., Robertson, G.J. and Reed, A., 2000. Common Eider (Somateria mollissima). In: A.

Poole and F. Gill (Editors), The Birds of North America, No. 546. The Birds of North

America, Inc., Philadelphia, pp. 1-32.

Grant, G.S., 1982. Avian incubation: egg temperature, nest humidity, and behavioral

thermoregulation in a hot environment. Ornithological Monographs 30: 1-75.

Grant, G.S., Paganelli, C.V., Pettit, T.N., Whittow, G.C. and Rahn, H., 1982. Determination of

fresh egg mass during natural incubation. Condor 84: 121-122.

Grant, G.S., Paganelli, C.V. and Rahn, H., 1984. Microclimate of Gull-billed Tern and Black

Skimmer nests. Condor 86: 337-338.

Grant, G.S., Pettit, T.N., Rahn, H., Whittow, G.C. and Paganelli, C.V., 1982a. Regulation of

water-loss from Bonin Petrel (Pterodroma hypoleuca) eggs. Auk 99: 236-242.

Grant, G.S., Pettit, T.N., Rahn, H., Whittow, G.C. and Paganelli, C.V., 1982b. Water loss from

Laysan and Black-footed Albatross eggs. Physiological Zoology 55: 405-414.

Greer, A.E., 1970. Evolutionary and systematic significance of crocodilian nesting habits. Nature

227: 523-524.

Grellet-Tinner, G., Chiappe, L., Norell, M. and Bottjer, D., 2006. Dinosaur eggs and nesting

behaviors: a paleobiological investigation. Palaeogeography, Palaeoclimatology,

Palaeoecology 232: 294-321.

Grellet-Tinner, G., Chiappe, L.M. and Coria, R., 2004. Eggs of titanosaurid sauropods from the

Upper Cretaceous of Auca Mahuevo (Argentina). Canadian Journal of Earth Sciences 41:

949-960.

Grellet-Tinner, G. and Fiorelli, L.E., 2010. A new Argentinean nesting site showing neosauropod 390

dinosaur reproduction in a Cretaceous hydrothermal environment. Nature

Communications 1: 10.1038/Ncomms1031.

Grellet-Tinner, G., Fiorelli, L.E. and Salvador, R.B., 2012. Water vapor conductance of the Lower

Cretaceous dinosaurian eggs from Sanagasta, La Rioja, Argentina: paleobiological and

paleoecological implications for South American faveoloolithid and megaloolithid eggs.

Palaios 27: 35-47.

Grellet-Tinner, G. and Makovicky, P., 2006. A possible egg of the dromaeosaur Deinonychus

antirrhopus: phylogenetic and biological implications. Canadian Journal of Earth Sciences

43: 705-719.

Griffiths, C.S., Barrowclough, G.F., Groth, J.G. and Mertz, L.A., 2007. Phylogeny, diversity, and

classification of the Accipitridae based on DNA sequences of the RAG-1 exon. Journal of

Avian Biology 38: 587-602.

Grigg, G. and Beard, L., 1985. Water loss and gain by eggs of Crocodylus porosus, related to

incubation age and fertility. In: G. Grigg, R. Shine and H. Ehmann (Editors), Biology of

Australasian Frogs and Reptiles. Royal Zoological Society of New South Wales, Sydney,

Australia, pp. 353-359.

Grigorescu, D., Garcia, G., Csiki, Z., Codrea, V. and Bojar, A.V., 2010. Uppermost Cretaceous

megaloolithid eggs from the Hateg Basin, Romania, associated with hadrosaur hatchlings:

search for explanation. Palaeogeography Palaeoclimatology Palaeoecology 293: 360-374.

Grigorescu, D., Weishampel, D.B., Norman, D., Seclamen, M., Rusu, M., Baltres, A. and

Teodorescu, V., 1994. Late Maastrichtian dinosaur eggs from the Hateg Basin (Romania).

In: K. Carpenter, K.F. Hirsch and J.R. Horner (Editors), Dinosaur Eggs and Babies.

Cambridge University Press, Cambridge, UK, pp. 75-87.

Groebbles, F., 1927. Oologische Studien. I. Mitt. Untersuchungen über Gewichte, spezifische 391

Gewichte und Zusammensetzung von Vogeleiern verschiedener Bebrütungsstadien.

Journal of Ornithology 75: 225-235.

Groebbles, F. and Mobert, F., 1927. Oologische studien. II. Mitteilupg. Ueber die kunstiiche

bestimmung der brutdauer einiger vogelarten mit hesonderser berucksichtigung des

eigewichts. Journal of Ornithology 17: 198-205.

Groombridge, B., 1982. The IUCN Amphibia - Reptilia Red Data Book Part 1: Testudines,

Crocodylia, Rhynchocephalia. IUCN, Gland, Switzerland, 426 pp.

Gross, A.O., 1938. Eider ducks of Kent's Islands. Auk 55: 387-4000.

Gullion, G.W., 1954. The reproductive cycle of American Coots in California. Auk 71: 366-412.

Hackett, S.J., Kimball, R.T., Reddy, S., Bowie, R.C.K., Braun, E.L., Braun, M.J., Chojnowski, J.L.,

Cox, W.A., Han, K.L., Harshman, J., Huddleston, C.J., Marks, B.D., Miglia, K.J., Moore,

W.S., Sheldon, F.H., Steadman, D.W., Witt, C.C. and Yuri, T., 2008. A phylogenomic study

of birds reveals their evolutionary history. Science 320: 1763-1768.

Hagelund, K. and Norderhaug, M., 1973. Studies of population changes and breeding processes in

a colony of Eiders (Somateria mollissima (L.)) in Svalbard. Norsk Polarinst. Årbok 1973:

141-161.

Hall, M.I., Kamilar, J.M. and Kirk, E.C., 2012. Eye shape and the nocturnal bottleneck of

mammals. Proceedings of the Royal Society B 279: 4962-4968.

Hall, P.M. and Johnson, D.R., 1987. Nesting biology of Crocodylus novaeguineae in Murray

District, Papua New Guinea. Herpetologica 43: 249-258.

Hamidu, J.A., Fasenko, G.M., Feddes, J.J.R., O'Dea, E.E., Ouellette, C.A., Wineland, M.J. and

Christensent, V.L., 2007. The effect of breeder genetic strain and parent flock age

on eggshell conductance and embryonic metabolism. Poultry Science 86: 2420-2432.

Hamilton, W.J., 1943. Nesting of the Eastern . Auk 60: 91-94. 392

Handrich, Y., 1989. Incubation water loss in King Penguin egg. I. Change in egg and brood pouch

parameters. Physiological Zoology 62: 96-118.

Hansell, M.H., 2000. Bird Nests and Construction Behaviour. Cambridge University Press,

Cambridge, UK, 280 pp.

Hansell, M.H. and Deeming, D.C., 2002. Location, structure and function of incubation sites. In:

D.C. Deeming (Editor), Avian Incubation: Behaviour, Environment, and Evolution.

Oxford University Press, Oxford, UK, pp. 8-27.

Harris, M.P., 1964. Aspects of the breeding biology of the gulls. Ibis 106: 432-456.

Harris, M.P., 1966. Breeding biology of the Manx Shearwater Puffinus puffinis. Ibis 108: 17-33.

Harris, R.B., Birks, S.M. and Leache, A.D., 2014. Incubator birds: biogeographical origins and

evolution of underground nesting in megapodes (Galliformes: Megapodiidae). Journal of

Biogeography 41: 2045-2056.

Harrison, C., 1975. A Field Guide to the Nests, Eggs and Nestlings of European Birds: with North

Africa and the Middle East. Collins, London, UK, 432 pp.

Hatch, J.J., 2002. Arctic Tern (Sterna paradisaea). In: A. Poole and F. Gill (Editors), The Birds of

North America, No. 707. The Birds of North America, Inc., Philadelphia, pp. 1-40.

Hatch, J.J., Brown, K.M., Hogan, G.G. and Morris, R.D., 2000. Great (Phalacrocorax

carbo). In: A. Poole and F. Gill (Editors), The Birds of North America, No. 553. The Birds

of North America, Inc., Philadelphia, pp. 1-32.

Hatch, J.J. and Weseloh, D.V., 1999. Double-crested Cormorant (Phalacrocorax auritus). In: A.

Poole and F. Gill (Editors), The Birds of North America, No. 441. The Birds of North

America, Inc., Philadelphia, pp. 1-38.

Hatch, S.A. and Nettleship, D.N., 1998. Northern Fulmar (Fulmarus glacialis). In: A. Poole and F.

Gill (Editors), The Birds of North America, No. 361. The Birds of North America, Inc., 393

Philadelphia, pp. 1-32.

Hayman, P. and Hume, R., 2007. Bird: the Ultimate Illustrated Guide to the Birds of Britain and

Europe. Octopus Publishing, London, UK, 552 pp.

Hechenleitner, E.M., Fiorelli, L.E., Grellet-Tinner, G., Leuzinger, L., Basilici, G., Taborda, J.R.A.,

De la Vega, S.R. and Bustamante, C.A., 2016. A new Upper Cretaceous titanosaur nesting

site from La Rioja (NW Argentina), with implications for titanosaur nesting strategies.

Palaeontology 59: 433-446.

Hechenleitner, E.M., Grellet-Tinner, G., Foley, M., Fiorelli, L.E. and Thompson, M.B., 2016.

Micro-CT scan reveals an unexpected high-volume and interconnected pore network in a

Cretaceous Sanagasta dinosaur eggshell. Journal of the Royal Society Interface 13:

10.1098/Rsif.2016.0008.

Heij, C.J., Rompas, C.F.E. and Moeliker, C.W., 1997. The biology of the Moluccan Megapode

Eulipoa wallacei (Aves, Galliformes, Megapodiidae) on Haruku and other Moluccan

Islands; part 2: final report. Deinsea 3: 1-123.

Heinroth, O., 1922. Die Beziehungen zwischen Vogelgewicht, Eigewicht, Gelegegewicht und

Brutdaner. Journal of Ornithology 70: 172-275.

Hepp, G.R. and Bellrose, F.C., 1995. (Aix sponsa). In: A. Poole and F. Gill (Editors),

The Birds of North America, No. 169. The Academy of Natural Sciences, Philadelphia,

and The American Ornithologists' Union, Washington, D.C., pp. 1-24.

Herman, A.B., Spicer, R.A. and Spicer, T.E.V., 2016. Environmental constraints on terrestrial

vertebrate behaviour and reproduction in the high Arctic of the Late Cretaceous.

Palaeogeography Palaeoclimatology Palaeoecology 441: 317-338.

Herron, J.C., Emmons, L.H. and Cadle, J.E., 1990. Observations on reproduction in the Black

Caiman, Melanosuchus niger. Journal of Herpetology 24: 314-316. 394

Higgins, P.J., 1999. Handbook of Australian, New Zealand and Antarctic Birds. Vol. 4: to

Dollarbird. Oxford University Press, Melbourne, Australia, 1248 pp.

Higgins, P.J. and Davies, S.J.J.F., 1996. Handbook of Australian, New Zealand and Antarctic Birds.

Vol. 3: Snipe to Pigeons. Oxford University Press, Melbourne, Australia, 1086 pp.

Higgins, P.J. and Peter, J.M., 2002. Handbook of Australian, New Zealand and Antarctic Birds.

Vol. 6: Pardalotes to Shrike-thrushes. Oxford University Press, Oxford, UK, 1262 pp.

Higgins, P.J., Peter, J.M. and Cowling, S.J., 2006. Handbook of Australian, New Zealand and

Antarctic Birds. Vol. 7: Boatbill to . Oxford University Press, Melbourne,

Australia, 1992 pp.

Higgins, P.J., Peter, J.M. and Steele, W.K., 2001. Handbook of Australian, New Zealand and

Antarctic Birds. Vol. 5: Tyrant-flycatchers to Chats. Oxford University Press, Melbourne,

Australia, 1272 pp.

Hipfner, J.M. and Chapdelaine, G., 2002. Razorbill (Alca torda). In: A. Poole and F. Gill (Editors),

The Birds of North America, No. 635. The Academy of Natural Sciences, Philadelphia,

and The American Ornithologists' Unnion, Washington, D.C., pp. 1-36.

Hipfner, J.M., Charleton, K. and Davies, W.E., 2004. Rates and consequences of relaying in

Cassin's Auklets Ptychoramphus aleuticus and Rhinoceros Auklets Cerorhinea

monocerata breeding in a seasonal environment. Journal of Avian Biology 35: 224-236.

Hirsch, K.F., 1994. Upper Jurassic eggshells from the Western Interior of North America. In: K.

Carpenter, K.F. Hirsch and J.R. Horner (Editors), Dinosaur Eggs and Babies. Cambridge

University Press, Cambridge, UK, pp. 137-150.

Hirsch, K.F. and Quinn, B., 1990. Eggs and eggshell fragments from the Upper Cretaceous Two

Medicine Formation of Montana. Journal of Vertebrate Paleontology 10: 491-511.

Hirsch, K.F., Stadtman, K.L., Miller, W.E. and Madsen, J.H., 1989. Upper Jurassic dinosaur egg 395

from Utah. Science 243: 1711-1713.

Hobson, K.A., 1997. (Phalacrocorax pelagicus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 282. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp. 1-28.

Hohman, W.L. and Lee, S.A., 2001. Fulvous Whistling-Duck (Dendrocygna bicolor). In: A. Poole

and F. Gill (Editors), The Birds of North America, No. 562. The Birds of North America,

Inc., Philadelphia, pp. 1-24.

Horner, J.R., 1982. Evidence of colonial nesting and 'site fidelity' among ornithischian dinosaurs.

Nature 297: 675-676.

Horner, J.R., 1999. Egg clutches and embryos of two hadrosaurian dinosaurs. Journal of Vertebrate

Paleontology 19: 607-611.

Horner, J.R., 2000. Dinosaur reproduction and parenting. Annual Review of Earth and Planetary

Sciences 28: 19-45.

Horner, J.R. and Currie, P.J., 1994. Embryonic and neonatal morphology and ontogeny of a new

species of Hypacrosaurus (Ornithischia, Lambeosauridae) from Montana and Alberta. In:

K. Carpenter, K.F. Hirsch and J.R. Horner (Editors), Dinosaur Eggs and Babies.

Cambridge University Press, Cambridge, UK, pp. 312-336.

Horner, J.R. and Makela, R., 1979. Nest of juveniles provides evidence of family structure among

dinosaurs. Nature 282: 296-298.

Horner, J.R., Padian, K. and de Ricqles, A., 2001. Comparative osteohistology of some embryonic

and perinatal archosaurs: developmental and behavioral implications for dinosaurs.

Paleobiology 27: 39-58.

Horton, D.H., 1932. The loss of weight of duck eggs during incubation. Poultry Science 11: 23-27.

Howell, T.R., 1979. Breeding Biology of the Egyptian Plover, Pluvianus aegyptinus. University of 396

California Publications in Zoology, 113. University of California Press, Berkeley, 76 pp.

Howell, T.R., Araya, B. and Millie, W.R., 1974. Breeding biology of the Gray Gull, Larus

modestus. University of California Publications in Zoology 104: 1-67.

Howell, T.R. and Bartholomew, G.A., 1961. Temperature regulation in nesting Bonin Island

Petrels, Wedge-tailed Shearwaters, and Christmas Island Shearwaters. Auk 78: 343-354.

Hoyt, D.F., 1979. Practical methods of estimating volume and fresh weight of bird eggs. Auk 96:

73-77.

Hoyt, D.F., 1980. Adaptation of avian eggs to incubation period: variability around allometric

regressions is correlated with time. American Zoologist 20: 417-425.

Hoyt, D.F., Board, R.G., Rahn, H. and Paganelli, C.V., 1979. The eggs of the Anatidae:

conductance, pore structure, and metabolism. Physiological Zoology 52: 438-450.

Hoyt, D.F. and Rahn, H., 1980. Respiration of avian embryos: a comparative analysis. Respiration

Physiology 39: 255-264.

Huber, B.T., Norris, R.D. and MacLeod, K.G., 2002. Deep-sea paleotemperature record of extreme

warmth during the Cretaceous. Geology 30: 123-126.

Huh, M., Kim, B.S., Woo, Y., Simon, D.J., Paik, I.S. and Kim, H.J., 2014. First record of a

complete giant theropod egg clutch from Upper Cretaceous deposits, South Korea.

Historical Biology 26: 218-228.

Huh, M. and Zelenitsky, D.K., 2002. Rich dinosaur nesting site from the Cretaceous of Bosung

County, Chullanam-do Province, South Korea. Journal of Vertebrate Paleontology 22:

716-718.

Hunt, H., 1973. Breeding Morelet's crocodile Crocodylus moreletii at Atlanta Zoo. International

Zoo Yearbook 13: 103-105.

Hunt, R.H., 1975. Maternal behavior in Morelet's Crocodile, Crocodylus moreleti. Copeia 4: 397

763-764.

Huntington, C.E., Butler, R.G. and Mauck, R.A., 1996. Leach's Storm-Petrel (Oceanodroma

leucorhoa). In: A. Poole and F. Gill (Editors), The Birds of North America, No. 233. The

Academy of Natural Sciences, Philadelphia, and The American Ornithologists' Union,

Washington, D.C., pp. 1-32.

Hussell, D.J.T., 1972. Factors affecting clutch size in Arctic . Ecological Monographs

42: 317-364.

Hussell, D.J.T. and Montgomerie, R., 2002. Lapland Longspur (Calcarius lapponicus). In: A.

Poole and F. Gill (Editors), The Birds of North America, No. 656. The Birds of North

America, Inc., Philadelphia, pp. 1-32.

Hussell, D.J.T. and Page, G.W., 1976. Observations on breeding biology of Black-bellied Plovers

on Devon-Island, NWT, Canada. Wilson Bulletin 88: 632-653.

Hutton, J.M., 1987. Incubation temperatures, sex ratios and sex determination in a population of

Nile crocodiles (Crocodylus niloticus). Journal of Zoology 211: 143-155.

Huynen, L., Gill, B.J., Millar, C.D. and Lambert, D.M., 2010. Ancient DNA reveals extreme egg

morphology and nesting behavior in New Zealand's extinct moa. Proceedings of the

National Academy of Sciences of the United States of America 107: 16201-16206.

Indrawan, M., Masala, Y. and Pesik, L., 1998. Observations on the breeding behaviour of Sula

Scrubfowl Megapodius bernsteinii in the Banggai Islands, Sulawesi, Indonesia. Forktail

13: 13-16.

Jackson, F.D., 2007. Titanosaur reproductive biology: comparison of the Auca Mahuevo titanosaur

nesting locality (Argentina), to the Pinyes Megaloolithus nesting locality (Spain). Ph.D.

Thesis, Montana State University, Bozeman, 166 pp.

Jackson, F.D., Horner, J.R. and Varricchio, D.J., 2010. A study of a Troodon egg containing 398

embryonic remains using epifluorescence microscopy and other techniques. Cretaceous

Research 31: 255-262.

Jackson, F.D., Schaff, R.J., Varricchio, D.J. and Schmitt, J.G., 2015. A theropod nesting trace with

eggs from the Upper Cretaceous (Campanian) of Montana.

Palaios 30: 362-372.

Jackson, F.D., Schmitt, J.G. and Oser, S.E., 2013. Influence of Vertisol development on sauropod

egg taphonomy and distribution at the Auca Mahuevo locality, Patagonia, Argentina.

Palaeogeography Palaeoclimatology Palaeoecology 386: 300-307.

Jackson, F.D. and Varricchio, D.J., 2010. Fossil eggs and eggshell from the lowermost Two

Medicine Formation of western Montana, Sevenmile Hill locality. Journal of Vertebrate

Paleontology 30: 1142-1156.

Jackson, F.D. and Varricchio, D.J., 2016. Fossil egg and eggshells from the Upper Cretaceous Hell

Creek Formation, Montana. Journal of Vertebrate Paleontology 36: e1185432.

Jackson, F.D., Varricchio, D.J., Jackson, R.A., Vila, B. and Chiappe, L.M., 2008. Comparison of

water vapor conductance in a titanosaur egg from the Upper Cretaceous of Argentina and a

Megaloolithus siruguei egg from Spain. Paleobiology 34: 229-246.

Jaeckle, W.B., Kiefer, M., Childs, B., Harper, R.G., Rivers, J.W. and Peer, B.D., 2012. Comparison

of eggshell porosity and estimated gas flux between the brown-headed cowbird and two

common hosts. Journal of Avian Biology 43: 486-490.

James, J.D. and Thompson, J.E., 2001. Black-bellied Whistling-Duck (Dendrocygna autumnalis).

In: A. Poole and F. Gill (Editors), The Birds of North America, No. 578. The Birds of North

America, Inc., Philadelphia, pp. 1-20.

Jardine, D.R., 1981. First successful captive propagation of Schneider's Smooth-fronted Caiman,

Paleosuchus trigonatus. Herpetological Review 12: 58-60. 399

Jarvis, E.D., Mirarab, S., Aberer, A.J., Li, B., Houde, P., Li, C., Ho, S.Y.W., Faircloth, B.C.,

Nabholz, B., Howard, J.T., Suh, A., Weber, C.C., da Fonseca, R.R., Li, J.W., Zhang, F., Li,

H., Zhou, L., Narula, N., Liu, L., Ganapathy, G., Boussau, B., Bayzid, M.S., Zavidovych,

V., Subramanian, S., Gabaldon, T., Capella-Gutierrez, S., Huerta-Cepas, J., Rekepalli, B.,

Munch, K., Schierup, M., Lindow, B., Warren, W.C., Ray, D., Green, R.E., Bruford, M.W.,

Zhan, X.J., Dixon, A., Li, S.B., Li, N., Huang, Y.H., Derryberry, E.P., Bertelsen, M.F.,

Sheldon, F.H., Brumfield, R.T., Mello, C.V., Lovell, P.V., Wirthlin, M., Schneider, M.P.C.,

Prosdocimi, F., Samaniego, J.A., Velazquez, A.M.V., Alfaro-Nunez, A., Campos, P.F.,

Petersen, B., Sicheritz-Ponten, T., Pas, A., Bailey, T., Scofield, P., Bunce, M., Lambert,

D.M., Zhou, Q., Perelman, P., Driskell, A.C., Shapiro, B., Xiong, Z.J., Zeng, Y.L., Liu, S.P.,

Li, Z.Y., Liu, B.H., Wu, K., Xiao, J., Yinqi, X., Zheng, Q.M., Zhang, Y., Yang, H.M., Wang,

J., Smeds, L., Rheindt, F.E., Braun, M., Fjeldsa, J., Orlando, L., Barker, F.K., Jonsson, K.A.,

Johnson, W., Koepfli, K.P., O'Brien, S., Haussler, D., Ryder, O.A., Rahbek, C., Willerslev,

E., Graves, G.R., Glenn, T.C., McCormack, J., Burt, D., Ellegren, H., Alstrom, P., Edwards,

S.V., Stamatakis, A., Mindell, D.P., Cracraft, J., Braun, E.L., Warnow, T., Jun, W., Gilbert,

M.T.P. and Zhang, G.J., 2014. Whole-genome analyses resolve early branches in the tree of

life of modern birds. Science 346: 1320-1331.

Jenkyns, H.C., Forster, A., Schouten, S. and Damste, J.S.S., 2004. High temperatures in the Late

Cretaceous Arctic Ocean. Nature 432: 888-892.

Jensen, A.L. and Kjelgaard-Hansen, M., 2006. Method comparison in the clinical laboratory.

Veternary Clinical Pathology 35: 276–286.

Jensen, J.A., 1966. Dinosaur eggs from the Upper Cretaceous North Horn Formation of Central

Utah. Brigham Young University Geology Studies 13: 55-67.

Ji, Q., 2009. Study on dinosaur eggs in China: yesterday and today. Acta Geoscientia Sinica 30: 400

285-290.

Jiang, X.J., Liu, Y.Q., Ji, S.A., Zhang, X.L., Xu, L., Jia, S.H., Lü, J.C., Yuan, C.X. and Li, M., 2011.

Dinosaur-bearing strata and K/T boundary in the Luanchuan-Tantou Basin of western

Henan Province, China. Science China-Earth Sciences 54: 1-7.

Jin, X.S., Azuma, Y., Jackson, F.D. and Varricchio, D.J., 2007. Giant dinosaur eggs from the

Tiantai basin, Zhejiang Province, China. Canadian Journal of Earth Sciences 44: 81-88.

Jin, X.S., Jackson, F.D., Varricchio, D.J., Azuma, Y. and He, T., 2010. The first Dictyoolithus egg

clutches from the Lishui Basin, Zhejiang Province, China. Journal of Vertebrate

Paleontology 30: 188-195.

Joanen, T., 1969. Nesting ecology of alligators in Louisiana. Proceedings of the Annual

Conference, Southeastern Association of Game and Fish Commissioneres 24: 141-151.

Joanen, T. and McNease, L.L., 1989. Ecology and physiology of nesting and early development of

the American Alligator. American Zoologist 29: 987-998.

Johnson, K.P., De Kort, S., Dinwoodey, K., Mateman, A.C., Ten Cate, C., Lessells, C.M. and

Clayton, D.H., 2001. A molecular phylogeny of the dove genera Streptopella and Columba.

Auk 118: 874-887.

Johnson, R.E., Hendricks, P., Pattie, D.L. and Hunter, K.B., 2000. Brown-capped Rosy-finch

(Leucosticte australis). In: A. Poole and F. Gill (Editors), The Birds of North America, No.

536. The Birds of North America, Inc., Philadelphia, pp. 1-24.

Johnston, R.F., 1992. Rock Dove. In: A. Poole, P. Stettenheim and F. Gill (Editors), The Birds of

North America, No. 13. The Academy of Natural Sciences, Philadelphia, and The

American Ornithologists' Union, Washington, D.C., pp. 1-16.

Jones, D.N., 1988. Construction and maintenance of the incubation mounds of the Australian

Brush-Turkey Alectura lathami. Emu 88: 210-218. 401

Jones, D.N., Dekker, R.W.R.J. and Roselaar, C.S., 1995. The Megapodes. Oxford University Press,

Oxford, UK, 262 pp.

Jonsson, K.A. and Fjeldsa, J., 2006. A phylogenetic supertree of oscine birds (Aves:

Passeri). Zoologica Scripta 35: 149-186.

Kamal, I., 2002. Heermann's Gull (Larus heermanni). In: A. Poole and F. Gill (Editors), The Birds

of North America, No. 643. The Birds of North America, Inc., Philadelphia, pp. 1-16.

Kannan, R. and James, D.A., 1998. Common Peafowl (Pavo cristatus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 377. The Birds of North America, Inc.,

Philadelphia, pp. 1-16.

Kathju, K., 2007. Observations of unusual clutch size, renesting and egg concealment by Sarus

Cranes antigone in Gujarat, India. Forktail 23: 165-167.

Kear, J., 2005. Ducks, Geese and Swans. Oxford University Press, Oxford, UK, 832 pp.

Kellers, H.C., 1931. Biological collecting on "tin-can island". In: Explorations and Field-Work of

the Smithsonian Institution in 1931. Smithsonian Institution, Washington, pp. 67-76.

Kemper, J., Underhill, L.G., Roux, J.-P., Bartlett, P.A., Chesselet, Y.J., James, J.A.C., Jones, R.,

Uhongora, N.-N. and Wepener, S., 2007. Breeding patterns and factors influencing

breeding success of African Penguins Spheniscus demersus in Namibia. In: S.P. Kirkman

(Editor), Final Report of the BCLME (Benguela Current Large Marine Ecosystem) Project

on Top Predators as Biological Indicators of Ecosystem Change in the BCLME. Avian

Demography Unit, Cape Town, South Africa, pp. 89-99.

Kendeigh, S.C., 1940. Factors affecting length of incubation. Auk 57: 499-513.

Kendeigh, S.C., 1963. Thermodynamics of incubation in the House Wren, Troglodytes aedon.

Proceedings of the International Ornithological Congress 13: 884-904.

Kennedy, M. and Page, R.D.M., 2002. supertrees: combining partial estimates of 402

procellariiform phylogeny. Auk 119: 88-108.

Kennedy, M. and Spencer, H.G., 2000. Phylogeny, biogeography, and taxonomy of Australasian

teals. Auk 117: 154-163.

Kennedy, M. and Spencer, H.G., 2004. Phylogenies of the (Fregatidae) and

Tropicbirds (Phaethonidae), two divergent groups of the traditional order Pelecaniformes,

inferred from mitochondrial DNA sequences. Molecular Phylogenetics and Evolution 31:

31-38.

Kern, M.D., Cowie, R.J. and Yeager, M., 1992. Water loss, conductance, and structure of eggs of

Pied Flycatchers during egg laying and incubation. Physiological Zoology 65: 1162-1187.

Kern, M.D., Sogge, M.K. and Vanriper, C., 1990. Water-vapor pressure in nests of the San-Miguel

Island Song Sparrow. Condor 92: 761-767.

Khosla, A. and Sahni, A., 1995. Parataxonomic classification of Late Cretaceous dinosaur

eggshells from India. Journal of The Palaeontological Society of India 40: 87-102.

Kim, J.Y., Yang, S.Y., Choi, H., Seo, S.J. and Kim, K.S., 2011. Dinosaur eggs from the Cretaceous

Goseong Formation of Tongyeong City, southern coast of Korea. Journal of

Paleontological Society of Korea 27: 13-26.

Kim, S.B., Kim, Y.G., Jo, H.R., Jeong, K.S. and Chough, S.K., 2009. Depositional ,

architecture and environments of the Sihwa Formation (Lower Cretaceous), mid-west

Korea with special reference to dinosaur eggs. Cretaceous Research 30: 100-126.

Kitching, J.W., 1979. Preliminary report on a clutch of six dinosaurian eggs from the Upper

Triassic Elliot Formation, Northern Orange Free State. Palaeontologia Africana 22: 41-45.

Klima, J. and Jehl, J.J.R., 1998. Stilt (Calidris himantopus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 341. The Birds of North America, Inc.,

Philadelphia, pp. 1-20. 403

Kofron, C.P., 1989. Nesting ecology of the Nile crocodile (Crocodylus niloticuls). African Journal

of Ecology 27: 335-341.

Krebs, E.A. and Putland, D.A., 2004. Chic chicks: the evolution of chick ornamentation in rails.

Behavioral Ecology 15: 946-951.

Krumenacker, L.J., Simon, D.J., Scofield, G. and Varricchio, D.J., 2016. Theropod dinosaurs from

the - of eastern Idaho. Historical Biology:

10.1080/08912963.2015.1137913.

Ksepka, D.T., Bertelli, S. and Giannini, N.P., 2006. The phylogeny of the living and fossil

Sphenisciformes (penguins). Cladistics 22: 412-441.

Kuhk, R., 1949. Aus der Fortpflanzungsbiologie des Rauhfusskauzes, Aegolius funereus (L.). In: E.

Mayr and E. Schuz (Editors), Ornithologie als Biologische Wissenschaft. Carl Winter

Universitatsverlag, Heidelberg, Germany, pp. 171-182.

Kundrat, M., Cruickshank, A.R.I., Manning, T.W. and Nudds, J., 2008. Embryos of

therizinosauroid theropods from the Upper Cretaceous of China: diagnosis and analysis of

ossification patterns. Acta Zoologica 89: 231-251.

Kurochkin, E.N., Chatterjee, S. and Mikhailov, K.E., 2013. An embryonic Enantiornithine bird

and associated eggs from the Cretaceous of Mongolia. Paleontological Journal 47:

1252-1269.

Kurzanov, S.M. and Mikhailov, K.E., 1989. Dinosaur eggshells from the Lower Cretaceous of

Mongolia. In: D.D. Gillette and M.G. Lockley (Editors), Dinosaur Tracks and Traces.

Cambridge University Press, New York, pp. 109-113.

Kushlan, J.A. and Bildstein, K.L., 1992. White Ibis. In: A. Poole, P. Stettenheim and F. Gill

(Editors), The Birds of North America, No. 9. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp. 1-20. 404

Kushlan, J.A. and Hancock, A.H., 2005. Herons. Oxford University Press, Oxford, UK, 456 pp.

Kushlan, J.A. and Mazzotti, F.J., 1989. Population biology of the American Crocodile. Journal of

Herpetology 23: 7-21.

Kutner, M.H., Nachtsheim, C.J., Neter, J. and Li, W., 2005. Applied Linear Statistical Models.

McGraw-Hill Irwin, New York, 1396 pp.

Lack, D., 1968. Ecological Adaptations for Breeding in Birds. Methuen and Co. Ltd., London, UK,

409 pp.

Lance, V.A., 2003. Alligator physiology and life history: the importance of temperature.

Experimental Gerontology 38: 801-805.

Lang, J.W., Andrews, H. and Whitaker, R., 1989. Sex determination and sex ratios in Crocodylus

palustris. American Zoologist 29: 935-952.

Lang, J.W. and Andrews, H.V., 1994. Temperature-dependent sex determination in crocodilians.

Journal of Experimental Zoology 270: 28-44.

Lank, D.B., Oring, L.W. and Maxson, S.J., 1985. Mate and nutrient limitation of egg-laying in a

polyandrous shorebird. Ecology 66: 1513-1524.

Lavin, S.R., Karasov, W.H., Ives, A.R., Middleton, K.M. and Garland, T., 2008. Morphometrics of

the avian small intestine compared with that of nonflying mammals: a phylogenetic

approach. Physiological and Biochemical Zoology 81: 526-550.

Le Maho, Y., 1977. The : a strategy to live and breed in the cold: morphology,

physiology, ecology, and behavior distinguish the polar emperor penguin from other

penguin species, particularly from its close relative, the king penguin. American Scientist

65: 680-693.

Le Souef, D., 1899. On the habits of the mound-building birds of Australia. Ibis 5: 9-19.

Lee, D.S. and Haney, J.C., 1996. Manx Shearwater (Puffinus puffinus). In: A. Poole and F. Gill 405

(Editors), The Birds of North America, No. 257. The Academy of Natural Sciences,

Philadelphia, and the American Ornithologists' Union, Washington, D.C., pp. 1-28.

Lee, Y.-N., 2003. Dinosaur bones and eggs in South Korea. Memoir of the Fukui Prefectural

Dinosaur Museum 2: 113-121.

Leon-Velarde, F., Whittemburry, J., Carey, C. and Monge, C., 1984. Permeability of eggshells of

native in the Peruvian Andes. In: R.S. Seymour (Editor), Respiration and

Metabolism of Embryonic Vertebrates. Dr. W. Junk Publications, Dordrecht, Netherlands,

pp. 245-258.

Lerner, H.R.L. and Mindell, D.P., 2005. Phylogeny of eagles, Old World vultures, and other

Accipitridae based on nuclear and mitochondrial DNA. Molecular Phylogenetics and

Evolution 37: 327-346.

Lewis, F., 1940. Notes on the breeding habits of the Mallee-Fowl. Emu 40: 97-110.

Li, Y., Yin, Z. and Liu, Y., 1995. The discovery of a new genus of dinosaur egg from Xixia, Henan,

China. Journal of Wuhan Institute of Chemical Technology 17: 38-40.

Liang, X., Wen, S., Yang, D., Zhou, S. and Wu, S., 2009. Dinosaur eggs and dinosaur egg-bearing

deposits (Upper Cretaceous) of Henan Province, China: occurrences, palaeoenvironments,

taphonomy and preservation. Progress in Natural Science 19: 1587-1601.

Lill, A., 1979. Nest inattentiveness and its influence on development of the young in the superb

lyrebird. Condor 81: 225-231.

Lill, A., 1987. Water vapor conductance and shell characteristics of eggs.

Australian Journal of Zoology 35: 553-558.

Lill, A., 1990. Water vapor flux in the eggs of two species of rail (Rallidae) during incubation.

Proceedings of the Royal Society of Victoria 102: 67-70.

Lill, A. and Fell, P.J., 2007. Microclimate of nesting burrows of the Rainbow Bee-eater. Emu 107: 406

108-114.

Lincoln, G.A., 1974. Predation of incubator birds (Megapodius freycinet) by Komodo dragons

(Varanus komodoensis). Journal of Zoology 174: 419-428.

Liston, J.J. and Chapman, S.D., 2014. Alfred Nicholson Leeds and the first fossil egg attributed to

a 'saurian'. Historical Biology 26: 229-235.

Liu, J.Y., Wang, Q., Zhao, Z., Wang, X.L., Gao, C.L. and Shen, C.Z., 2013. A parataxonomic

revision of spheroolithid eggs from the Upper Cretaceous Quantou Formation in Changtu,

Liaoning. Vertebrata Palasiatica 51: 278-288.

Livezey, B.C., 1995. Phylogeny and evolutionary ecology of modern seaducks (Anatidae,

Mergini). Condor 97: 233-255.

Livezey, B.C., 1996. A phylogenetic analysis of modern pochards (Anatidae: Aythyini). Auk 113:

74-93.

Livezey, B.C., 2001. A phylogenetic analysis of modern sheldgeese and shelducks (Anatidae,

Tadornini). Ibis 139: 51-66.

Lomholt, J.P., 1976. Relationship of weight loss to ambient humidity of birds eggs during

incubation. Journal of Comparative Physiology 105: 189-196.

Lopez-Luna, M.A., Hidalgo-Mihart, M.G., Aguirre-Leon, G., Gonzalez-Ramon, M.D. and

Rangel-Mendoza, J.A., 2015. Effect of nesting environment on incubation temperature and

hatching success of Morelet's crocodile (Crocodylus moreletii) in an urban lake of

Southeastern Mexico. Journal of Thermal Biology 49/50: 66-73.

Lopez-Martinez, N., Moratalla, J.J. and Sanz, J.L., 2000. Dinosaurs nesting on tidal flats.

Palaeogeography Palaeoclimatology Palaeoecology 160: 153-163.

Lopez-Martinez, N. and Vicens, E., 2012. A new peculiar dinosaur egg, Sankofa pyrenaica oogen.

nov. oosp. nov. from the Upper Cretaceous coastal deposits of the Aren Formation, 407

south-central Pyrenees, Lleida, Catalonia, Spain. Palaeontology 55: 325-339.

Lowther, P.E. and Collins, C.T., 2002. Black Swift (Cypseloides niger). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 676. The Birds of North America, Inc.,

Philadelphia, pp. 1-16.

Lowther, P.E., Diamond, A.W., Kress, S.W., Robertson, G.J. and Russell, K., 2002.

(Fratercula arctica). In: A. Poole and F. Gill (Editors), The Birds of North America, No.

709. The Birds of North America, Inc., Philadelphia, pp. 1-24.

Lü, J., Azuma, Y., Huang, D., Noda, Y. and Qiu, L., 2006. New troodontid dinosaur eggs from the

Heyuan Basin of Guangdong Province, Southern China. In: J. Lü, Y. Kobayashi, D. Huang

and Y.N. Lee (Editors), Papers from the 2005 Heyuan International Dinosaur Symposium.

Geological Publishing House, Beijing, China, pp. 11-18.

Lutz, P.L. and Dunbar-Cooper, A., 1984. The nest environment of the American Crocodile

(Crocodylus acutus). Copeia 1984: 153-161.

Lyson, T.R. and Longrich, N.R., 2011. Spatial niche partitioning in dinosaurs from the latest

cretaceous (Maastrichtian) of North America. Proceedings of the Royal Society B 278:

1158-1164.

MacKinnon, J., 1978. Sulawesi Megapodes. Journal of the World Pheasant Association 3: 96-103.

MacLean, G.L., 1974. Egg-covering in the Charadrii. Ostrich 45: 167-174.

Maddison, W.P. and Maddison, D.R., 2010. Mesquite: a modular system for evolutionary analysis.

Version 2.73. See http://mesquiteproject.org.

Magnusson, W.E., 1979. Maintenance of temperature of crocodile nests (Reptilia, Crocodilidae).

Journal of Herpetology 13: 439-443.

Magnusson, W.E., Lima, A.P. and Sampaio, R.M., 1985. Sources of heat for nests of Paleosuchus

trigonatus and a review of crocodilian nest temperatures. Journal of Herpetology 19: 408

199-207.

Makovicky, P.J. and Grellet-Tinner, G., 2001. Association between theropod eggshell and a

specimen of Deinonychus antirrhopus. In: A.M. Bravo and T. Reyes (Editors), First

International Symposium on Dinosaur Eggs and Babies Extended Abstracts, Catalonia,

Spain, pp. 123-128.

Mallory, M. and Metz, K., 1999. Common Merganser (Mergus merganser). In: A. Poole and F.

Gill (Editors), The Birds of North America, No. 442. The Birds of North America, Inc.,

Philadelphia, pp. 1-28.

Manning, T.H., 1981. Analysis of weight lost by eggs of 11 species of birds during incubation.

Canadian Field-Naturalist 95: 63-68.

Manning, T.H., 1982. Daily measurements of variation in weight-loss of eggs of seven Passerine

species before and during natural incubation. Canadian Journal of Zoology 60: 3143-3149.

Manning, T.W., Joysey, K.A. and Cruickshank, A.R.I., 1997. Observations of microstructures

within dinosaur eggs from Henan Province, Peoples' Republic of China. In: D.L. Wolberg,

E. Stump and R.D. Rosenberg (Editors), Dinofest International: Proceedings of a

Symposium Held at Arizona State University. Academy of Natural Sciences, Philadelphia,

pp. 287-290.

Manuwal, D.A. and Thoresen, A.C., 1993. Cassin's Auklet (Ptychoramphus aleuticus). In: A.

Poole and F. Gill (Editors), The Birds of North America, No. 50. The Academy of Natural

Sciences, Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp.

1-20.

Marchant, S. and Higgins, P.J., 1990. Handbook of Australian, New Zealand and Antarctic Birds.

Vol. 1: Raptors to Ducks. Oxford University Press, Melbourne, Australia, 735 pp.

Marchant, S. and Higgins, P.J., 1994. Handbook of Australian, New Zealand and Antarctic Birds. 409

Vol. 2: Raptors to Lapwings, 2. Oxford University Press, Oxford, UK, 1048 pp.

Marti, C.D., 1992. Barn . In: A. Poole, P. Stettenheim and F. Gill (Editors), The Birds of North

America, No. 1. The Academy of Natural Sciences, Philadelphia, and The American

Ornithologists' Union, Washington, D.C., pp. 1-15.

Martin, T.E., 1993. Nest predation and nest sites: new perspectives on old patterns. Bioscience 43:

523-532.

Martin, T.E., 2002. A new view of avian life-history evolution tested on an incubation paradox.

Proceedings of the Royal Society B 269: 309-316.

Marzola, M., Russo, J. and Mateus, O., 2015. Identification and comparison of modern and fossil

crocodilian eggs and eggshell structures. Historical Biology 27: 115-133.

Massaro, M. and Davis, L.S., 2005. Differences in egg size, shell thickness, pore density, pore

diameter and water vapour conductance between first and second eggs of snares penguins

Eudyptes robustus and their influence on hatching asynchrony. Ibis 147: 251-258.

Mateus, I., Mateus, H., Antunes, M.T., Mateus, O., Taquet, P., Ribeiro, V. and Manuppella, G.,

1997. Clutches, eggs and embryos of therapod dinosaur in the upper Jurassic of Lourinha,

Portugal. Comptes Rendus de l'Académie des Sciences-Series II: A-Earth and Planetary

Sciences 325: 71-78.

Mateus, I., Mateus, H., Antunes, M.T., Mateus, O., Taquet, P., Ribeiro, V. and Manuppella, G.,

1998. Upper Jurassic theropod dinosaur embryos from Lourinha (Portugal). Memorias da

Academia de Ciencias de Lisboa 37: 101-109.

Mateus, O., Walen, A. and Antunes, M.T., 2006. The large theropod fauna of the Lourinha

Formation (Portugal) and its similarity to the Morrison Formation, with a description of a

new species of Allosaurus. New Mexico Museum of Natural History and Science Bulletin

36: Paleontology and Geology of the Upper Jurassic Morrison Formation: 1-7. 410

Maunder, J.E. and Threlfall, W., 1972. The breeding biology of the Black-legged Kittiwake in

Newfoundland. Auk 89: 789-816.

Mauro, I., 2005. Field discovery, mound characteristics, bare parts, vocalisations and behaviour of

Bruijn's Brush-turkey (Aepypodius bruijnii). Emu 105: 273-281.

Maxwell, W.D. and Horner, J.R., 1994. Neonate dinosaurian remains and dinosaurian eggshell

from the , Montana. Journal of Vertebrate Paleontology 14: 143-146.

Mayfield, H., 1961. Nesting success calculated from exposure. Wilson Bulletin 73: 255-261.

Mayfield, H.F., 1975. Suggestions for calculating nest success. Wilson Bulletin 87: 456-466.

Mayr, G., 2010. interrelationships: morphology and the new molecular phylogenies. Emu

110: 348-357.

Mazzotti, F.J., 1989. Factors affecting the nesting success of the American Crocodile, Crocodylus

acutus, in Florida Bay. Bulletin of Marine Science 44: 220-228.

McCartney, C., 1990. Breeding of West African Dwarf crocodiles (Osteolaemus tetraspis

tetraspis) at the Port Elizabeth Park. Journal of the Herpetological Association of

Africa 38: 44-48.

McCracken, K.G., Barger, C.P. and Sorenson, M.D., 2010. Phylogenetic and structural analysis of

the HbA (A/A) and HbD (D/A) hemoglobin genes in two high-altitude waterfowl from

the and the Andes: bar-headed goose (Anser indicus) and Andean goose

(Chloephaga melanoptera). Molecular Phylogenetics and Evolution 56: 649-658.

McIlhenny, E.A., 1934. Notes on incubation and growth of alligators. Copeia 2: 80-88.

McIntyre, J.W. and Barr, J.F., 1997. (Gavia immer). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 313. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Unnion, Washington, D.C., pp. 1-32.

McNicholl, M.K., Lowther, P.E. and Hall, J.A., 2001. Forster's Tern (Sterna forsteri). In: A. Poole 411

and F. Gill (Editors), The Birds of North America, No. 595. The Birds of North America,

Inc., Philadelphia, pp. 1-24.

Meanley, B., 1969. Natural History of the King Rail. United States Department of the Interior

Bureau of Sport Fisheries and Wildlife, Washington D.C., 108 pp.

Medem, F., 1958. The crocodilian genus Paleosuchus. Fieldiana Zoology 39: 227-247.

Medem, F., 1971. The reproduction of the Dwarf Caiman Paleosuchus palpebrosus, Crocodiles:

Proceedings of the 1st Working Meeting of Crocodile Specialists. IUCN, New York, pp.

159-165.

Medem, F., 1981. Los Crocodylia de sur America Volumen 1. Los Crocodylia de .

Universidad Nacional de Colombia, Bogota, Colombia, 354 pp.

Mehlum, F., 1991. Egg predation in a breeding colony of the Common Eider Somateria mollissima

in Kongsfjorden , Svalbard. Norsk Polarinstitutt Skrifter 195: 37-45.

Merne, O.J., 1974. Ducks, Geese and Swans. St. Martin's Press, New York, 160 pp.

Metzen, W.D., 1977. Nesting ecology of alligators on the Okefenokee National Wildlife Refuge.

Proceedings of the Southeastern Association of Fish and Wildlife Agencies 31: 29-32.

Mezquida, E.T., 2001. Ecologia reproductiva de un ensamble de aves del desierto del monte

central, Argentina. Ph.D. Thesis, Universidad Autonoma de Madrid, Madrid, Spain, 149

pp.

Mickey, F.W., 1943. Breeding habits of Mccown's Longspur. Auk 60: 181-209.

Midford, P.E., Garland Jr., T. and Maddison, W.P., 2010. PDAP package of Mesquite, version 1.15.

See http://mesquiteproject.org/pdap_mesquite/index.html.

Mikhailov, K., 1994a. Eggs of sauropod and ornithopod dinosaurs from the Cretaceous deposits of

Mongolia. Paleontological Journal 28: 141-159.

Mikhailov, K., 1994b. Theropod and protoceratopsian dinosaur eggs from the Cretaceous of 412

Mongolia and . Paleontological Journal 28: 101-120.

Mikhailov, K., Sabath, K. and Kurzanov, S., 1994. Eggs and nests from the Cretaceous of

Mongolia. In: K. Carpenter, K.F. Hirsch and J.R. Horner (Editors), Dinosaur Eggs and

Babies. Cambridge University Press, Cambridge, UK, pp. 88-115.

Mikhailov, K.E., 1991. Classification of fossil eggshells of amniotic vertebrates. Acta

Palaeontologica Polonica 36: 193-238.

Mikhailov, K.E., Bray, E.S. and Hirsch, K.E., 1996. Parataxonomy of fossil egg remains

(Veterovata): principles and applications. Journal of Vertebrate Paleontology 16: 763-769.

Miller, A.H., 1931. Systematic revision and natural history of the American Shrikes (Lanius).

University of California Publications in Zoology 38: 11-242.

Miranda, M.P., de Moraes, G.V., Martins, E.N., Maia, L.C.P. and Barbosa, O.R., 2002. Thermic

variation in incubation and development of pantanal Caiman (Caiman crocodilus yacare)

(Daudin, 1802) kept in metabolic box. Brazilian Archives of Biology and Technology 45:

333-342.

Mitchell, C.D., 1994. Trumpeter Swan (Cygnus buccinator). In: A. Poole and F. Gill (Editors), The

Birds of North America, No. 105. The Academy of Natural Sciences, Philadelphia, and

The American Ornithologists' Union, Washington, D.C., pp. 1-24.

Mitchell, K.J., Llamas, B., Soubrier, J., Rawlence, N.J., Worthy, T.H., Wood, J., Lee, M.S.Y. and

Cooper, A., 2014. Ancient DNA reveals elephant birds and kiwi are sister taxa and clarifies

ratite bird evolution. Science 344: 898-900.

Modha, M.L., 1967. The ecology of the Nile Crocodile (Crocodylus niloticus Laurenti) on Central

Island, Lake Rudolf. African Journal of Ecology 5: 74-95.

Mohabey, D.M., 1996. A new oospecies, Megaloolithus matleyi, from the

(Upper Cretaceous) of Chandrapur district, Maharashtra, India, and general remarks on the 413

palaeoenvironment and nesting behaviour of dinosaurs. Cretaceous Research 17: 183-196.

Mohabey, D.M., 1998. Systematics of Indian Upper Cretaceous dinosaur and chelonian eggshells.

Journal of Vertebrate Paleontology 18: 348-362.

Mohabey, D.M., 2000. Indian dinosaur eggs: a review. Journal of the Geological Society of India

58: 479-508.

Mohabey, D.M., 2001. Indian Upper Cretaceous (Maastrichtian) dinosaur eggs: their

parataxonomy and implications in understanding the nesting behaviour. In: A.M. Bravo

and T. Reyes (Editors), First International Symposium on Dinosaur Eggs and Babies

Extended Abstracts, Catalonia, Spain, pp. 139-154.

Mohabey, D.M. and Mathur, U.B., 1989. Upper Cretaceous dinosaur eggs from new localities of

Gujarat, India. Journal of the Geological Society of India 33: 32-37.

Mones, A., 1980. Nuevos elementos de la paleoherpetofauna del uruguay (Crocodilia y

Dinosauria). Actas del Segundo Congreso Argentino de Paleontología y Bioestratigrafía y

Primer Congreso Latinoamericano de Paleontología: Buenos Aires, Argentina, 2-6 de abril,

1978, 1: 265-278.

Montini, J.P., Piña, C.I., Larriera, A., Siroski, P. and Verdade, L.M., 2006. The relationship

between nesting habitat and hatching success in Caiman latirostris (Crocodylia,

Alligatoridae). Phyllomedusa: Journal of Herpetology 5: 91-96.

Moore, J.C., 1953. The crocodile in the Everglades National Park. Copeia 1953: 54-59.

Moorman, T.E. and Gray, P.N., 1994. Mottled Duck (Anas fulvigula). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 81. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp. 1-20.

Moreno-Azanza, M., Canudo, J.I. and Gasca, J.M., 2014a. Spheroolithid eggshells in the Lower

Cretaceous of Europe. Implications for eggshell evolution in ornithischian dinosaurs. 414

Cretaceous Research 51: 75-87.

Moreno-Azanza, M., Canudo, J.I. and Gasca, J.M., 2014b. Unusual theropod eggshells from the

Early Cretaceous of the Iberian Range, Spain. Acta Palaeontologica

Polonica 59: 843-854.

Morgan, F.D., 1976. Nesting studies of the (Passerina cyanea) at Thornhill,

Indiana. Proceedings of the Indiana Academy of Science 86: 461-465.

Morgan, K.R., Paganelli, C.V. and Rahn, H., 1978. Egg weight loss and nest humidity during

incubation in two Alaskan gulls. Condor 80: 272-275.

Morris, R.D. and Chardine, J.W., 1992. The breeding biology and aspects of the feeding ecology of

Brown Noddies Anous stolidus nesting near Culebra, Puerto Rico, 1985-1989. Journal of

Zoology 226: 65-79.

Moskoff, W. and Bevier, L.R., 2002. Mew Gull (Larus canus). In: A. Poole and F. Gill (Editors),

The Birds of North America, No. 687. The Birds of North America, Inc., Philadelphia, pp.

1-28.

Motani, R. and Schmitz, L., 2011. Phylogenetic versus functional signals in the evolution of

form-function relationships in terrestrial vision. Evolution 65: 2245-2257.

Mou, Y., 1992. Nest environments of the Late Cretaceous dinosaur eggs from Nanxiong Basin,

Guangdong Province. Vertebrata Palasiatica 30: 120-134.

Mowbray, T.B., 2002. Northern Gannet (Morus bassanus). In: A. Poole and F. Gill (Editors), The

Birds of North America, No. 693. The Birds of North America, Inc., Philadelphia, pp. 1-28.

Mowbray, T.B., Ely, C.R., Sedinger, J.S. and Trost, R.E., 2002. (Branta canadensis).

In: A. Poole and F. Gill (Editors), The Birds of North America, No. 682. The Birds of North

America, Inc., Philadelphia, pp. 1-44.

Muller, M.J. and Storer, R.W., 1999. Pied-billed Grebe (Podilymbus podiceps). In: A. Poole and F. 415

Gill (Editors), The Birds of North America, No. 410. The Birds of North America, Inc.,

Philadelphia, pp. 1-32.

Nair, T. and Katdare, S., 2013. Dry-season assessment of (Gavialis gangeticus) in the

Betwa, Ken and Son Rivers, India. In: World crocodile conference. Proceedings of the

22nd Working Meeting of the IUCN-SSC Crocodile Specialist Group. IUCN, Gland,

Switzerland, pp. 53-65.

Neil, W.T., 1946. Notes on Crocodylus novaeguineae. Copeia 1946: 17-20.

Nelson, J.B., 1966. Breeding biology of Gannet Sula bassana on Bass Rock Scotland. Ibis 108:

584-626.

Nelson, J.B., 1971. Biology of Abbotts Booby Sula abbotti. Ibis 113: 429-467.

Nelson, J.B., 2006. Pelicans, and Their Relatives. Oxford University Press, Oxford,

UK, 680 pp.

Nelson, S.K., 1997. Marbled Murrelet (Brachyramphus marmoratus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 276. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Unnion, Washington, D.C., pp. 1-32.

Nesbitt, S.J., 2011. The early evolution of archosaurs: relationships and the origin of major clades.

Bulletin of the American Museum of Natural History 352: 1-288.

Nisbet, I.C.T., 2002. (Sterna hirundo). In: A. Poole and F. Gill (Editors), The Birds

of North America, No. 618. The Birds of North America, Inc., Philadelphia, pp. 1-40.

Nolan, V.J., Ketterson, E.D., Cristol, D.A., Rogers, C.M., Clotfelter, E.D., Titus, R.C., Schoech,

S.J. and Snajdr, E., 2002. Dark-eyed Junco (Junco hyemalis). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 716. The Birds of North America, Inc.,

Philadelphia, pp. 1-44.

Norell, M., Clark, J.M. and Chiappe, L.M., 2001. An embryonic oviraptorid (Dinosauria: 416

Theropoda) from the Upper Cretaceous of Mongolia. American Museum Novitates 3315:

1-17.

Norell, M.A., Clark, J.M., Demberelyin, D., Rhinchen, B., Chiappe, L.M., Davidson, A.R.,

Mckenna, M.C., Altangerel, P. and Novacek, M.J., 1994. A theropod dinosaur embryo and

the affinities of the Flaming Cliffs dinosaur eggs. Science 266: 779-782.

Novas, F.E., Salgado, L., Suarez, M., Agnolin, F.L., Ezcurra, M.D., Chimento, N.R., de la Cruz, R.,

Isasi, M.P., Vargas, A.O. and Rubilar-Rogers, D., 2015. An enigmatic plant-eating

theropod from the Late Jurassic period of Chile. Nature 522: 331–334.

O'dea, E.E., Fasenko, G.M., Feddes, J.J.R., Robinson, F.E., Segura, J.C., Ouellette, C.A. and van

Middelkoop, J.H., 2004. Investigating the eggshell conductance and embryonic

metabolism of modern and unselected domestic avian genetic strains at two flock ages.

Poultry Science 83: 2059-2070.

Oaks, J.R., 2011. A time-calibrated species tree of Crocodylia reveals a recent radiation of the true

crocodiles. Evolution 65: 3285-3297.

Ogden, J.C., 1978. Status and nesting biology of American Crocodile, Crocodylus acutus,

(Reptilia, Crocodilidae) in Florida. Journal of Herpetology 12: 183-196.

Oliveira, C.E.M., Santucci, R.M., Andrade, M.B., Fulfaro, V.J., Basilio, J.A.F. and Benton, M.J.,

2011. Crocodylomorph eggs and eggshells from the (),

Upper Cretaceous of Brazil. Palaeontology 54: 309-321.

Ouboter, P.E. and Nanhoe, L.M.R., 1987. Notes on nesting and parental care in Caiman crocodilus

crocodilus in northern Suriname and an analysis of crocodilian nesting habitats.

Amphibia-Reptilia 8: 331-347.

Packard, G.C. and Packard, M.J., 1994. The physiological ecology of reptilian eggs and embryos.

Biology of the Reptilia 16: 525-605. 417

Packard, G.C., Taigen, T.L., Packard, M.J. and Shuman, R.D., 1979. Water vapor conductance of

testudinian and crocodilian eggs (class Reptilia). Respiration Physiology 38: 1-10.

Paganelli, C.V., 1991. The avian eggshell as a mediating barrier: respiratory gas fluxes and

pressures during development. In: D.C. Deeming and M.W.J. Ferguson (Editors), Egg

Incubation: Its Effects on Embryonic Development in Birds and Reptiles. Cambridge

University Press, Cambridge, UK, pp. 261-275.

Paganelli, C.V., Olszowka, A. and Ar, A., 1974. Avian egg: surface area, volume, and density.

Condor 76: 319-325.

Pagel, M.D., 1992. A method for the analysis of comparative data. Journal of Theoretical Biology

156: 431-442.

Paik, I.S., Huh, M. and Kim, H.J., 2004. Dinosaur egg-bearing deposits (Upper Cretaceous) of

Boseong, Korea: occurrence, palaeoenvironments, taphonomy, and preservation.

Palaeogeography Palaeoclimatology Palaeoecology 205: 155-168.

Paik, I.S., Kim, H.J. and Huh, M., 2012. Dinosaur egg deposits in the Cretaceous Gyeongsang

Supergroup, Korea: diversity and paleobiological implications. Journal of Asian Earth

Sciences 56: 135-146.

Parsons, K.C. and Master, T.L., 2000. Snowy Egret (Egretta thula). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 489. The Birds of North America, Inc.,

Philadelphia, pp. 1-24.

Passing, H. and Bablok, W., 1983. A new biometrical procedure for testing the equality of

measurements from two different analytical methods. Journal of Clinical Chemistry and

Clinical Biochemistry 21: 709–720.

Patterson, R.L., 1952. The Sage in Wyoming. Sage Books Wyoming Game and Fish

Commission, Denver, 341 pp. 418

Peitz, C., 2001. Megaloolithid dinosaur eggs from the Maastrichtian of Catalunya (NE-Spain) -

parataxonomic implications and stratigraphic utility. In: A.M. Bravo and T. Reyes (Editors),

First International Symposium on Dinosaur Eggs and Babies Extended Abstracts,

Catalonia, Spain, pp. 155-160.

Pettit, T.N., Grant, G.S., Whittow, G.C., Rahn, H. and Paganelli, C.V., 1981. Respiratory gas

exchange and growth of white tern embryos. Condor 83: 355-361.

Phillips, M.J., Gibb, G.C., Crimp, E.A. and Penny, D., 2010. Tinamous and moa flock together:

mitochondrial genome sequence analysis reveals independent losses of flight among ratites.

Systematic Biology 59: 90-107.

Piatt, J.F. and Kitaysky, A.S., 2002a. Horned Puffin (Fratercula corniculata). In: A. Poole and F.

Gill (Editors), The Birds of North America, No. 603. The Birds of North America, Inc.,

Philadelphia, pp. 1-28.

Piatt, J.F. and Kitaysky, A.S., 2002b. Tufted Puffin (Fratercula cirrhata). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 708. The Birds of North America, Inc.,

Philadelphia, pp. 1-32.

Pierotti, R.J. and Good, T.P., 1994. Herring Gull (Larus argentatus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 124. The Academy of Natural Sciences,

Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp. 1-28.

Pina, C.I., Larriera, A. and Cabrera, M.R., 2003. Effect of incubation temperature on incubation

period, sex ratio, hatching success, and survivorship in Caiman latirostris (Crocodylia,

Alligatoridae). Journal of Herpetology 37: 199-202.

Pizzey, G., 1980. A field Guide to the Birds of Australia. Collins, Sydney, Australia, 460 pp.

Platt, S.G., Rainwater, T.R., Thorbjarnarson, J.B. and McMurry, S.T., 2008. Reproductive

dynamics of a tropical freshwater crocodilian: Morelet's crocodile in northern Belize. 419

Journal of Zoology 275: 177-189.

Platt, S.G. and Thorbjarnarson, J.B., 2000. Nesting ecology of the American crocodile in the

coastal zone of Belize. Copeia: 869-873.

Plumb, W.J., 1965. Observations on the breeding biology of the Razorbill. British Birds 58:

449-456.

Pooley, A.C., 1969. Preliminary studies on the breeding of the Nile Crocodile Crocodylus niloticus,

in Zululand. Lammergeyer 10: 22-44.

Post, W.J., Poston, J.P. and Bancroft, G.T., 1996. Boat-tailed Grackle (Quiscalus major). In: A.

Poole and F. Gill (Editors), The Birds of North America, No. 207. The Academy of Natural

Sciences, Philadelphia, and The American Ornithologists' Union, Washington, D.C., pp.

1-28.

Prokop, P. and Trnka, A., 2011. Why do grebes cover their nests? Laboratory and field tests of two

alternative hypotheses. Journal of Ethology 29: 17-22.

Proudfoot, G.A., Gehlbach, F.R. and Honeycutt, R.L., 2007. Mitochondrial DNA variation and

phylogeography of the Eastern and Western screech-. Condor 109: 617-627.

Qian, M., Jiang, Y., Chen, R., Jiang, Y., Zhang, Y.L. and Xin, G., 2008. Late Cretaceous troodontid

eggs from Tiantai, Zhejiang. Acta Palaeontologica Sinica: 248-255.

Rahn, H., 1984. Factors controlling the rate of incubation water loss in bird eggs. In: R.S. Seymour

(Editor), Respiration and Metabolism of Embryonic Vertebrates. Dr. W. Junk Publications,

Dordrecht, Netherlands, pp. 271-288.

Rahn, H., Ackerman, R.A. and Paganelli, C.V., 1977. Humidity in avian nest and egg water loss

during incubation. Physiological Zoology 50: 269-283.

Rahn, H. and Ar, A., 1974. Avian egg: incubation time and water loss. Condor 76: 147-152.

Rahn, H., Carey, C., Balmas, K., Bhatia, B. and Paganelli, C., 1977. Reduction of pore area of 420

avian eggshell as an adaptation to altitude. Proceedings of the National Academy of

Sciences of the United States of America 74: 3095-3098.

Rahn, H. and Dawson, W.R., 1979. Incubation water loss in eggs of Heermann's and Western Gulls.

Physiological Zoology 52: 451-460.

Rahn, H., Greene, D.G., Toien, O., Krog, J. and Mehlum, F., 1984. Estimated laying dates and

eggshell conductance of the Fulmar and Brunnich's Murre in Spitsbergen. Ornis

Scandinavica 15: 110-114.

Rahn, H. and Hammel, H.T., 1982. Incubation water loss, shell conductance, and pore dimensions

in Adelie Penguin eggs. Polar Biology 1: 91-97.

Rahn, H. and Huntington, C.E., 1988. Eggs of Leach's Storm Petrel: O2 uptake, water loss, and

microclimate of the nest. Comparative Biochemistry and Physiology A 91: 519-521.

Rahn, H., Krog, J. and Mehlum, F., 1983. Microclimate of the nest and egg water loss of the Eider

Somateria mollissima and other waterfowl in Spitsbergen. Polar Research 1: 171-184.

Rahn, H. and Paganelli, C.V., 1990. Gas fluxes in avian eggs: driving forces and the pathway for

exchange. Comparative Biochemistry and Physiology A 95: 1-15.

Rahn, H., Paganelli, C.V. and Ar, A., 1987. Pores and gas exchange of avian eggs: a review.

Journal of Experimental Zoology: 165-172.

Rahn, H., Paganelli, C.V., Nisbet, I.C.T. and Whittow, G.C., 1976. Regulation of incubation water

loss in eggs of seven species of terns. Physiological Zoology 49: 245-259.

Rahn, H. and Whittow, G.C., 1988. Adaptations to a pelagic life: eggs of the Albatross, Shearwater

and Petrel. Comparative Biochemistry and Physiology A 91: 415-423.

Ramirez, J.L., Miyaki, C.Y. and Del Lama, S.N., 2013. Molecular phylogeny of

Thereskiornithidae (Aves: Pelecaniformes) based on nuclear and mitochondrial DNA.

Genetics and Molecular Research 12: 2740-2750. 421

Rand, A.L. and Gilliard, E.T., 1967. Handbook of New Guinea Birds. Weidenfeld and Nicolson,

London, UK, 612 pp.

Rao, R.J., 1988. Nesting Ecology of the Gharial in National Chambal Sanctuary. Wildlife Institute

of India, Dehradun, India, 105 pp.

Rao, R.J. and Gaurjwar, R.K., 2013. Crocodile human conflict in National Chambal Sanctuary,

India. In: World Crocodile Conference: Proceedings of the 22nd Working Meeting of the

IUCN-SSC Crocodile Specialist Group. IUCN, Gland, Switzerland, pp. 105-109.

Rao, R.J., Tagor, S., Singh, H. and Dasgupta, N., 2013. Monitoring of Gharial (Gavialis

gangeticus) and its habitat in the National Chambal Sanctuary, India. In: World Crocodile

Conference: Proceedings of the 22nd Working Meeting of the IUCN-SSC Crocodile

Specialist Group. IUCN, Gland, Switzerland, pp. 66-73.

Raynor, G.S. and Wilcox, L., 1980. Observations on the life-history of Willets on Long-Island,

New-York. Wilson Bulletin 92: 253-258.

Reisz, R.R., Evans, D.C., Roberts, E.M., Sues, H.D. and Yates, A.M., 2012. Oldest known

dinosaurian nesting site and reproductive biology of the sauropodomorph

Massospondylus. Proceedings of the National Academy of Sciences of the United States of

America 109: 2428-2433.

Reisz, R.R., Huang, T.D., Roberts, E.M., Peng, S.R., Sullivan, C., Stein, K., LeBlanc, A.R.H.,

Shieh, D., Chang, R.S., Chiang, C.C., Yang, C.W. and Zhong, S.M., 2013. Embryology of

Early Jurassic dinosaur from China with evidence of preserved organic remains. Nature

496: 210-214.

Reisz, R.R., Scott, D., Sues, H.D., Evans, D.C. and Raath, M.A., 2005. Embryos of an Early

Jurassic prosauropod dinosaur and their evolutionary significance. Science 309: 761-764.

Retallack, G.J., 2001. Soils of the Past: an Introduction to Paleopedology. Blackwell Science Ltd., 422

London, UK, 404 pp.

Ribeiro, C.M.M., 2002. Ovo e fragmentos de cascas de ovos de dinossauros, provenientes da

regiao de Peiropolis, Uberaba, Minas Gerais. Arquivos do Museu Nacional 60: 223-228.

Ribeiro, V., Mateus, O., Holwerda, F., Araujo, R. and Castanhinha, R., 2014. Two new theropod

egg sites from the Late Jurassic Lourinha Formation, Portugal. Historical Biology 26:

206-217.

Ricklefs, R.E., 1993. Sibling competition, hatching asynchrony, incubation period, and lifespan in

altricial birds. Current Biology 11: 199-276.

Ricklefs, R.E. and Starck, J.M., 1998. The evolution of the developmental mode in birds. In: J.M.

Starck and R.E. Ricklefs (Editors), Avian Growth and Development: Evolution within the

Altricial-Precocial Spectrum. Oxford University Press, New York, pp. 366-380.

Rigby, J.K., Snee, L.W., Unruh, D.M., Harland, S.S., , J., Li, F., Rigby, J.K. and Kowalis, B.J.,

1993. 40Ar/39Ar and U-Pb dates for dinosaur extinction, Nanxiong Basin, Guangdong

Province, People's Republic of China. Geological Society of America Abstracts with

Program 25: A296.

Riley, J. and Huchzermeyer, F.W., 1999. African dwarf crocodiles in the Likouala swamp forests

of the congo basin: habitat, density, and nesting. Copeia: 313-320.

Ripley, S.D., 1960. Distribution and niche differentiation in species of megapodes in the Moluccas

and western Papuan Area. Proceedings of the International Ornithological Congress 12:

631-640.

Rivas, J.A., Aktay, S.A. and Owens, R.Y., 2001. Paleosuchus trigonatus (Schneider's

Smooth-fronted Caiman) nesting. Herpetological Review 32: 251.

Robertson, G.J. and Savard, J.-P.L., 2002. Long-tailed Duck (Clangula hyemalis). In: A. Poole and

F. Gill (Editors), The Birds of North America, No. 651. The Birds of North America, Inc., 423

Philadelphia, pp. 1-28.

Robertson, H.A. and James, P.C., 1988. Morphology and egg measurements of seabirds breeding

on Great Salvage Island, North Atlantic. British Ornithologists' Club 108: 79-87.

Rogers, K.C., Whitney, M., D'Emic, M. and Bagley, B., 2016. Precocity in a tiny titanosaur from

the Cretaceous of Madagascar. Science 352: 450-453.

Rokitka, M.A. and Rahn, H., 1987. Regional differences in shell conductance and pore density of

avian eggs. Respiration Physiology 68: 371-376.

Roper, D.S., 1983. Egg incubation and laying behavior of the incubator bird Megapodius freycinet

on Savo. Ibis 125: 384-389.

Roudybush, T., Hoffman, L. and Rahn, H., 1980. Conductance, pore geometry, and water loss of

eggs of Cassin's Auklet. Condor 82: 105-106.

Ruxton, G.D., Birchard, G.F. and Deeming, D.C., 2014. Incubation time as an important influence

on egg production and distribution into clutches for sauropod dinosaurs. Paleobiology 40:

323-330.

Sabath, K., 1991. Upper Cretaceous amniotic eggs from the Gobi Desert. Palaeontologica 36:

151-192.

Sahni, A., Tandon, S.K., Jolly, A., Bajpai, S., Sood, A. and Srinivasan, S., 1994. Upper Cretaceous

dinosaur eggs and nesting sites from the Deccan volcano: sedimentary province of

Peninsular India. In: K. Carpenter, K.F. Hirsch and J.R. Horner (Editors), Dinosaur Eggs

and Babies. Cambridge University Press, Cambridge, UK, pp. 204-226.

Salgado, L., Coria, R.A., Ribeiro, C.M.M., Garrido, A., Rogers, R., Simon, M.E., Arcucci, A.B.,

Rogers, K.C., Carabajal, A.P., Apesteguia, S., Fernandez, M., Garcia, R.A. and Talevi, M.,

2007. Upper Cretaceous dinosaur nesting sites of Rio Negro (Salitral Ojo de Agua and

Salinas de Trapalco-Salitral de Santa Rosa), northern Patagonia, Argentina. Cretaceous 424

Research 28: 392-404.

Salgado, L., Ribeiro, C.M., Garcia, R.A. and Fernandez, M.S., 2009. Late Cretaceous

megaloolithid eggs from Salitral de Santa Rosa (Rio Negro, Patagonia, Argentina):

inferences on the titanosaurian reproductive biology. Ameghiniana 46: 605-620.

Sander, P.M., Peitz, C., Gallemi, J. and Cousin, R., 1998. Dinosaurs nesting on a red beach?

Comptes Rendus de l'Académie des Sciences-Series II: A-Earth and Planetary Sciences

327: 67-74.

Sander, P.M., Peitz, C., Jackson, F.D. and Chiappe, L.M., 2008. Upper Cretaceous titanosaur

nesting sites and their implications for sauropod dinosaur reproductive biology.

Palaeontographica Abteilung A: Palaozoologie-Stratigraphie 284: 69-107.

Sankaran, R., 1995. The distribution, status and conservation of the Nicobar Megapode

Megapodius nicobariensis. Biological Conservation 72: 17-25.

Sanz, J.L., Moratalla, J.J., Diazmolina, M., Lopezmartinez, N., Kalin, O. and Vianeyliaud, M.,

1995. Dinosaur nests at the sea shore. Nature 376: 731-732.

Sato, T., Cheng, Y.N., Wu, X.C., Zelenitsky, D.K. and Hsiao, Y.F., 2005. A pair of shelled eggs

inside a female dinosaur. Science 308: 375-375.

Schmitz, L. and Motani, R., 2011. Nocturnality in dinosaurs inferred from scleral ring and orbit

morphology. Science 332: 705-708.

Schonwetter, M., 1960–1967. Handbuch der Oologie. Akademie Verlag, Berlin, Germany, 929 pp.

Schrantz, F.G., 1943. Nest life of the Eastern Yellow Warbler. Auk 60: 367-387.

Schreiber, E.A., Feare, C.J., Harrington, B.A., Murray, J., B.G., Robertson, J., W.B., Robertson,

M.J. and Woolfenden, G.E., 2002. Sooty Tern (Sterna fuscata). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 665. The Birds of North America, Inc.,

Philadelphia, pp. 1-32. 425

Seddon, P.J. and van Heezik, Y., 1991. Effects of hatching order, sbling asymmetries, and nest site

on survival analysis of Jackass Penguin chicks. Auk 108: 548-555.

Seijas, A.E., 1998. The Orinoco Crocodile (Crocodylus intemedius) in the Cojedes River System,

Venezuela: population status and ecological characteristics. Ph.D. Thesis, University of

Florida, Gainesville, 191 pp.

Sellés, A.G., Bravo, A.M., Delclos, X., Colombo, F., Marti, X., Ortega-Blanco, J., Parellada, C.

and Galobart, A., 2013. Dinosaur eggs in the Upper Cretaceous of the Coll de Nargo area,

Lleida Province, south-central Pyrenees, Spain: oodiversity, biostratigraphy and their

implications. Cretaceous Research 40: 10-20.

Sellés, A.G. and Galobart, A., 2016. Reassessing the endemic European Upper Cretaceous

dinosaur egg Cairanoolithus. Historical Biology 28: 583-596.

Sellés, A.G., Via, B. and Galobart, A., 2014. Spheroolithus europaeus, oosp. nov. (late

Maastrichtian, Catalonia), the youngest oological record of hadrosauroids in Eurasia.

Journal of Vertebrate Paleontology 34: 725-729.

Sellés, A.G., Vila, B. and Galobart, A., 2014. Diversity of theropod ootaxa and its implications for

the latest Cretaceous dinosaur turnover in southwestern Europe. Cretaceous Research 49:

45-54.

Serventy, D.L. and Whittell, H.M., 1962. Birds of Western Australia. Paterson Brokensha, Perth,

Australia, 427 pp.

Seymour, R.S., 1979. Dinosaur eggs: gas conductance through the shell, water loss during

incubation and clutch size. Paleobiology 5: 1-11.

Seymour, R.S. and Ackerman, R.A., 1980. Adaptations to underground nesting in birds and

reptiles. American Zoologist 20: 437-447.

Seymour, R.S. and Bradford, D.F., 1992. Temperature regulation in the incubation mounds of the 426

Australian Brush-Turkey. Condor 94: 134-150.

Seymour, R.S. and Rahn, H., 1978. Gas conductance in the eggshell of the mound-building Brush

Turkey. In: J. Piiper (Editor), Respiratory Function in Birds, Adult and Embryonic.

Springer-Verlag, Berlin, Germany, pp. 243-246.

Seymour, R.S. and Visschedijk, A.H.J., 1988. Effects of variation in total and regional shell

conductance on air cell gas tensions and regional gas exchange in eggs. Journal of

Comparative Physiology B 158: 229-236.

Seymour, R.S., Vleck, D. and Vleck, C.M., 1986. Gas exchange in the incubation mounds of

megapode birds. Journal of Comparative Physiology B 156: 773-782.

Seymour, R.S., Vleck, D., Vleck, C.M. and Booth, D.T., 1987. Water relations of buried eggs of

mound building birds. Journal of Comparative Physiology B 157: 413-422.

Sheldon, F.H., 1987. Phylogeny of herons estimated from DNA-DNA hybridization data. Auk

104: 97-108.

Sheldon, F.H., Jones, C.E. and McCracken, K.G., 2000. Relative patterns and rates of evolution in

heron nuclear and mitochondrial DNA. Molecular Biology and Evolution 17: 437-450.

Sheldon, F.H. and Slikas, B., 1997. Advances in ciconiiform systematics 1976-1996. Colonial

Waterbirds 20: 106-114.

Sherman, A.R., 1911. Nest life of the Screech Owl. Auk 28: 155-168.

Sherman, A.R., 1913. The nest life of the Sparrow Hawk. Auk 30: 406-418.

Sibal, M.C., Sarsagat, I.G. and Satake, Y., 1992. Captive breeding of C. mindorensis and C.

porosus. In: Proceedings of the Workshop on the Prospects and Future Strategy of

Crocodile Conservation of the Two Species (Crocodylus mindorensis and Crocodylus

porosus) Occurring in the Philippines, pp. 36-44.

Silyn-Roberts, H., 1983. The pore geometry and structure of the eggshell of the North Island 427

Brown Kiwi, Apteryx australis mantelli. Journal of Microscopy 130: 23-36.

Sinclair, J.R., 2001. Temperature regulation in mounds of three sympatric species of megapode

(Aves: Megapodiidae) in Papua New Guinea: testing the 'Seymour Model'. Australian

Journal of Zoology 49: 675-694.

Sivakumar, K. and Sankaran, R., 2003. The incubation mound and hatching success of the Nicobar

Megapode Megapodius nicobariensis Blyth. Journal of the Bombay Natural History

Society 100: 375-387.

Skeel, M.A. and Mallory, E.P., 1996. Whimbrel (Numenius phaeopus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 219. The Academy of Natural Sciences,

Philadelphia, and the American Ornithologists' Union, Washington, D.C., pp. 1-28.

Skutch, A.F., 1985. Clutch size, nesting success, and predation on nests of neotropical birds,

reviewed. Ornithological Monographs 36: 575-594.

Smith, N.D., 2010. Phylogenetic analysis of Pelecaniformes (Aves) based on csteological data:

implications for waterbird phylogeny and fossil calibration studies. PLoS One 5:

10.1371/journal.pone.0013354.

Smith, T., Codrea, V.A., Sasaran, E., Van Itterbeeck, J., Bultynck, P., Csiki, Z., Dica, P., Farcas, C.,

Folie, A., Garcia, G. and Godefroit, P., 2002. A new exceptional vertebrate site from the

Late Cretaceous of the Hateg Basin (Romania). Studia Universitatis Babes-Bolyai,

Geologia, Special Issue 1: 321-330.

Snow, B., 1960. The breeding biology of the Shag Phalacrocorax aristotelis on the Island of

Lundy, Bristol Channel. Ibis 102: 554-575.

Snow, D.W. and Perrins, C.M., 1998. The Birds of the Western Palearctic, Vol. 1. Non-Passerines.

Oxford University Press, Oxford, UK, 1008 pp.

Snyder, G.K., Black, C.P. and Birchard, G.F., 1982. Development and metabolism during hypoxia 428

in embryos of high altitude Anser Indicus versus sea level Branta canadensis geese.

Physiological Zoology 55: 113-123.

Sochava, A.V., 1969. Dinosaur eggs from the Upper Cretaceous of the Gobi Desert.

Paleontological Journal 4: 517-527.

Sokal, R.R. and Rohlf, F.J., 1995. Biometry: the Principles and Practice of Statistics in Biological

Research. W.H. Freeman and Company, New York, 880 pp.

Somaweera, R. and Shine, R., 2013. Nest-site selection by crocodiles at a rocky site in the

Australian tropics: making the best of a bad lot. Austral Ecology 38: 313-325.

Sotherland, P.R., Ashen, M.D., Shuman, R.D. and Tracy, C.R., 1984. The water balance of bird

eggs incubated in water. Physiological Zoology 57: 338-348.

Soto, M., Perea, D. and Cambiaso, A., 2012. First sauropod (Dinosauria: ) remains from

the Guichon Formation, Late Cretaceous of Uruguay. Journal of South American Earth

Sciences 33: 68-79.

Srivastava, A.K. and Mankar, R.S., 2015. Megaloolithus dinosaur nest from the Lameta Formation

of Salbardi area, districts Amravati, Maharashtra and Betul, Madhya Pradesh. Journal of

the Geological Society of India 85: 457-462.

Srivastava, S., Mohabey, D.M., Sahni, A. and Pant, S.C., 1986. Upper Cretaceous dinosaur egg

clutches from Kheda District (Gujarat, India). Palaeontographica Abteilung A:

Palaozoologie-Stratigraphie 193: 219-233.

Staniewicz, A., 2011. Diet and demography of Tomistoma schlegelii in Mesangat Lake, East

Kalimantan, Indonesia: June–September 2011. University of Bristol, School of Biological

Sciences. Unpublished Report for EAZA IUCN/SSC Campaign: See

http://www.dght-zgap-artenschutz.de/media/projekte/mesangat/2011%20Mesangat%20To

mistoma%20report.pdf. 429

Starck, J.M. and Ricklefs, R.E., 1998. Patterns of development: the altricial-precocial spectrum.

In: J.M. Starck and R.E. Ricklefs (Editors), Avian Growth and Development: Evolution

within the Altricial-Precocial Spectrum. Oxford University Press, New York, pp. 3-30.

Staton, M.A. and Dixon, J.R., 1977. Breeding biology of the Spectacled Caiman, Caiman

crocodilus crocodilus, in the Venezuelan Llanos. In: Wildlife Research Report 5. United

States Department of the Interior, Fish and Wildlife Service, Washington, pp. 1-21.

Steen, J.B., Andersen, O., Saebo, A., Pedersen, H.C. and Erikstad, K.E., 1988. Viability of newly

hatched chicks of Lagopus l. lagopus. Ornis Scandinavica 19: 93-96.

Steen, J.B. and Unander, S., 1985. Breeding biology of the Svalbard Rock Ptarmigan Lagopus

mutus hyperboreus. Ornis Scandinavica 16: 191-197.

Stempniewicz, L., 1981. Breeding biology of the Little Auk Plautus alle in the Hornsund region,

Spitsbergen. Acta Ornithologica 18: 1-26.

Stieglit, W.O. and Wilson, C.T., 1968. Breeding biology of Florida Duck. Journal of Wildlife

Management 32: 921-934.

Stinson, D.W. and Glass, P.O., 1992. The Micronesian megapode Megapodius laperouse:

conservation and research needs. Zoologische Verhandelingen 278: 53-55.

Stoker, C., Zayas, M.A., Ferreira, M.A., Durando, M., Galoppo, G.H., Rodriguez, H.A., Repetti,

M.R., Beldomenico, H.R., Caldini, E.G., Luque, E.H. and Munoz-de-Toro, M., 2013. The

eggshell features and clutch viability of the Broad-snouted Caiman (Caiman latirostris) are

associated with the egg burden of organochlorine compounds. Ecotoxicology and

Environmental Safety 98: 191-195.

Stonehouse, B., 1962. The tropic birds (Genus Phaethon) of Ascension Island. Ibis 103: 124-161.

Storer, R.W. and Nuechterlein, G.L., 1992. Western and Clark's Grebe. In: A. Poole, P. Stettenheim

and F. Gill (Editors), The Birds of North America, No. 26. The Academy of Natural 430

Sciences, Philadelphia, and The American Ornithologists' Unnion, Washington, D.C., pp.

1-24.

Stresemann, E., 1914. Beiträge zur Kenntnis der Avifauna von . Notitates Zoologicae 21:

358-400.

Stuebing, R. and Johan, Z., 1986. The megapodes of Pulau Tiga. The Sabah Museum and State

Archives Journal 1: 16-49.

Sumner, E.L., 1928. Notes on the development of young Screech Owls. Condor 30: 333-338.

Sumner, E.L., 1929a. Notes on the growth and behavior of young Golden Eagles. Auk 46:

161-169.

Sumner, F.L., 1929b. Comparative studies in the growth of young raptors. Condor 31: 85-111.

Suzuki, S. and Watabe, M., 2000a. Report on the Japan - Mongolia Joint Paleontological

Expedition to the Gobi desert, 1995. Hayashibara Museum of Natural Sciences Research

Bulletin 1: 45-57.

Suzuki, S. and Watabe, M., 2000b. Report on the Japan - Mongolia Joint Paleontological

Expedition to the Gobi desert, 1998. Hayashibara Museum of Natural Sciences Research

Bulletin 1: 83-98.

Swanepoel, D.G.J., Ferguson, N.S. and Perrin, M.R., 2000. Nesting ecology of Nile crocodiles

(Crocodylus niloticus) in the Olifants River, Kruger National Park. Koedoe 43: 35-46.

Swanson, D.L. and Garland, T., 2009. The evolution of high summit metabolism and cold

tolerance in birds and its impact on present-day distributions. Evolution 63: 184-194.

Swart, D. and Rahn, H., 1988. Microclimate of ostrich nests: measurements of egg temperature

and nest humidity using egg hygrometers. Journal of Comparative Physiology B 157:

845-854.

Swart, D., Rahn, H. and Dekock, J., 1987. Nest microclimate and incubation water loss of eggs of 431

the African ostrich (Struthio camelus var. domesticus). Journal of Experimental Zoology:

239-246.

Sykes, P.W., 1987. Some aspects of the breeding biology of the Snail Kite in Florida. Journal of

Field Ornithology 58: 171-189.

Takashima, R., Nishi, H., Huber, B.T., and Leckie, R.M. 2006. Greenhouse world and the

Mesozoic Ocean. Oceanography 19: 82-92.

Tanaka, K., Lü, J., Liu, Y., Huang, Z., Kobayashi, Y., Huang, D. and Zelenitsky, D.K., 2012.

Statistical approach for classification of dinosaur eggs from the Heyuan Basin at the

northeast of Guangdong Provice. Acta Geologica Sinica 86: 294-303.

Tanaka, K., Lü, J.C., Kobayashi, Y., Zelenitsky, D.K., Xu, L., Jia, S.G., Qin, S.A. and Tang, M.N.,

2011. Description and phylogenetic position of dinosaur eggshells from the Luanchuan

area of western Henan Province, China. Acta Geologica Sinica (English Edition) 85:

66-74.

Tanaka, K. and Zelenitsky, D.K., 2014a. Comparisons between experimental and morphometric

water vapor conductance in the eggs of extant birds and crocodiles: implications for

predicting nest type in dinosaurs. Canadian Journal of Zoology 92: 1049-1058.

Tanaka, K. and Zelenitsky, D.K., 2014b. Relationships between nest humidity and nest types in

living archosaurs. Historical Biology 26: 122-131.

Tanaka, K., Zelenitsky, D.K., Saegusa, H., Ikeda, T., DeBuhr, C.L. and Therrien, F., 2016.

Dinosaur eggshell assemblage from Japan reveals unknown diversity of small theropods.

Cretaceous Research 57: 350-363.

Tanaka, K., Zelenitsky, D.K. and Therrien, F., 2015. Eggshell porosity provides insight on

evolution of nesting in dinosaurs. PLoS One 10: 10.1371/journal.pone.0142829.

Tanaka, K., Zelenitsky, D.K., Williamson, T., Weil, A. and Therrien, F., 2011. Fossil eggshells 432

from the Upper Cretaceous (Campanian) Fruitland Formation, New Mexico. Historical

Biology 23: 41-55.

Tarduno, J.A., Brinkman, D.B., Renne, P.R., Cottrell, R.D., Scher, H. and Castillo, P., 1998.

Evidence for extreme climatic warmth from Late Cretaceous Arctic vertebrates. Science

282: 2241-2244.

Temple, S.A., 2002. Dickcissel (Spiza americana). In: A. Poole and F. Gill (Editors), The Birds of

North America, No. 703. The Birds of North America, Inc., Philadelphia, pp. 1-24.

Thomas, G.H., Wills, M.A. and Szekely, T., 2004. A supertree approach to shorebird phylogeny.

BMC Evolutionary Biology 4: 10.1186/1471-2148-4-28.

Thompson, M.B. and Goldie, K.N., 1990. Conductance and structure of eggs of Adelie Penguins,

Pygoscelis adeliae, and its implications for incubation. Condor 92: 304-312.

Thompson, S.C. and Raveling, D.G., 1988. Nest insulation and incubation constancy of arctic

geese. Wildfowl 39: 124-132.

Thorbjarnarson, J., Wang, X.M. and He, L.J., 2001. Reproductive ecology of the Chinese alligator

(Alligator sinensis) and implications for conservation. Journal of Herpetology 35: 553-558.

Thorbjarnarson, J.B., 1996. Reproductive characteristics of the order Crocodylia. Herpetologica

52: 8-24.

Thorbjarnarson, J.B. and Hernandez, G., 1993. Reproductive ecology of the Orinoco Crocodile

(Crocodylus intermedius) in Venezuela. I. Nesting ecology and egg and clutch

relationships. Journal of Herpetology 27: 363-370.

Tickell, W.L.N., 1968. The biology of the Great Albatrosses, Diomedea exulans and Diomedea

epomophora. Antarctic Research Series 12: 1-55.

Titman, R.D., 1999. Red-breasted Merganser (Mergus serrator). In: A. Poole and F. Gill (Editors),

The Birds of North America, No. 443. The Birds of North America, Inc., Philadelphia, pp. 433

1-24.

Toien, O., Paganelli, C.V., Rahn, H. and Johnson, R.R., 1988. Diffusive resistance of avian

eggshell pores. Respiration Physiology 74: 345-354.

Tracy, D.M., Schamel, D. and Dale, J., 2002. Red Phalarope (Phalaropus fulicarius). In: A. Poole

and F. Gill (Editors), The Birds of North America, No. 698. The Birds of North America,

Inc., Philadelphia, pp. 1-32.

Tullett, S.G., 1976. The avian eggshell: a mediating boundary. Ph.D. Thesis, The University of

Bath, Bath, UK, 198 pp.

Tullett, S.G., 1981. Theoretical and practical aspects of eggshell porosity. Turkeys 29: 24-28.

Tullett, S.G. and Board, R.G., 1977. Determinants of avian eggshell porosity. Journal of Zoology

183: 203-211.

Tullett, S.G. and Deeming, D.C., 1982. The relationship between eggshell porosity and oxygen

consumption of the embryo in the domestic fowl. Comparative Biochemistry and

Physiology A 72: 529-533.

Turner, A.H., Makovicky, P.J. and Norell, M.A., 2012. A review of dromaeosaurid systematics and

paravian phylogeny. Bulletin of the American Museum of Natural History: 5-206.

Twedt, D.J. and Crawford, R.D., 1995. Yellow-headed Blackbird (Xanthocephalus

xanthocephalus). In: A. Poole and F. Gill (Editors), The Birds of North America, No. 192.

The Academy of Natural Sciences, Philadelphia, and The American Ornithologists'

Unnion, Washington, D.C., pp. 1-28.

Tyler, C. and Fowler, S., 1979. Size, shape and orientation of pore grooves in the egg shells of

Rhea sp. Journal of Zoology 187: 283-290.

Tyler, C. and Simkiss, K., 1959. A study of the egg shells of ratite birds. Proceedings of the

Zoological Society of London 133: 201-243. 434

Upchurch, G.R., Kiehl, J., Shields, C., Scherer, J. and Scotese, C., 2015. Latitudinal temperature

gradients and high-latitude temperatures during the latest Cretaceous: congruence of

geologic data and climate models. Geology 43: 683-686.

Uspenski, S.M., 1958. The bird Bazaars of Novaya Zemlya. In: Translations of Russian Game

Reports. Canadian Wildlife Service, Montreal, Canada, pp. 31-74. van Hinsbergen, D.J.J., de Groot, L.V., van Schaik, S.J., Spakman, W., Bijl, P.K., Sluijs, A.,

Langereis, C.G. and Brinkhuis, H., 2015. A paleolatitude calculator for paleoclimate

studies. PLoS One 10: 10.1371/journal.pone.0126946. van Straelen, V., 1925. The microstructure of the dinosaurian egg-shells from the Cretaceous beds

of Mongolia. American Museum Novitates 173: 1-4. van Weerd, M., 2010. Philippine Crocodile Crocodylus mindorensis. In: S.C. Manolis and C.

Stevenson (Editors), Crocodiles: Status, Survey and Conservation Action Plan. Crocodile

Specialist Group, Darwin, Australia, pp. 71-78. van Weerd, M., van der Ploeg, J., Rodriguez, D., Guerrero, J., Tarun, B., Telan, S. and de Jonge, J.,

2006. Philippine crocodile conservation in Northeast Luzon: an update of population status

and new insights into Crocodylus mindorensis ecology. In: Crocodiles: Proceedings of the

18th Working Meeting of the Crocodile Specialist Group. IUCN - the World Conservation

Union, Gland, Switzerland and Cambridge, UK, pp. 306-321.

Varricchio, D.J. and Berta, D.E., 2015. Revisiting Sabath's "larger avian eggs" from the Gobi

Cretaceous. Acta Palaeontologica Polonica 60: 11-25.

Varricchio, D.J., Horner, J.R. and Jackson, F.D., 2002. Embryos and eggs for the Cretaceous

theropod dinosaur Troodon formosus. Journal of Vertebrate Paleontology 22: 564-576.

Varricchio, D.J., Jackson, F., Borkowski, J.J. and Horner, J.R., 1997. Nest and egg clutches of the

dinosaur Troodon formosus and the evolution of avian reproductive traits. Nature 385: 435

247-250.

Varricchio, D.J., Jackson, F. and Trueman, C.N., 1999. A nesting trace with eggs for the

Cretaceous theropod dinosaur Troodon formosus. Journal of Vertebrate Paleontology 19:

91-100.

Varricchio, D.J. and Jackson, F.D., 2004. A phylogenetic assessment of prismatic dinosaur eggs

from the Cretaceous Two Medicine Formation of Montana. Journal of Vertebrate

Paleontology 24: 931-937.

Varricchio, D.J., Jackson, F.D., Jackson, R.A. and Zelenitsky, D.K., 2013. Porosity and water

vapor conductance of two Troodon formosus eggs: an assessment of incubation strategy in

a maniraptoran dinosaur. Paleobiology 39: 278-296.

Varricchio, D.J., Jin, X.-S. and Jackson, F.D., 2015. Lay, brood, repeat: nest reuse and site fidelity

in ecologic time for two Cretaceous troodontid dinosaurs. Journal of Vertebrate

Paleontology 35: e932797.

Varricchio, D.J., Moore, J.R., Erickson, G.M., Norell, M.A., Jackson, F.D. and Borkowski, J.J.,

2008. Avian had dinosaur origin. Science 322: 1826-1828.

Venczel, M., Vasile, S. and Csiki-Sava, Z., 2015. A Late Cretaceous madtsoiid snake from

Romania associated with a megaloolithid egg nest - paleoecological inferences. Cretaceous

Research 55: 152-163.

Verbeek, N.A.M., 1993. Glaucous-winged Gull (Larus glaucescens). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 59. The Academy of Natural Sciences,

Philadelphia, and the American Ornithologists' Union, Washington, D.C., pp. 1-20.

Vianey-Liaud, M., Hirsch, K.E., Sahni, A. and Sige, B., 1997. Late Cretaceous Peruvian eggshells

and their relationships with Laurasian and Eastern Gondowanian material. Geobios 30:

75-90. 436

Vianey-Liaud, M., Khosla, A. and Garcia, G., 2003. Relationships between European and Indian

dinosaur eggs and eggshells of the oofamily Megaloolithidae. Journal of Vertebrate

Paleontology 23: 575-585.

Vianey-Liaud, M. and Lopez-Martinez, N., 1997. Late Cretaceous dinosaur eggshells from the

Tremp Basin, southern Pyrenees, Lleida, Spain. Journal of Paleontology 71: 1157-1171.

Vianey-Liaud, M., Mallan, P., Buscail, O. and Montgelard, C., 1994. Review of French dinosaur

eggshells; morphology, structure, mineral, and organic composition. In: K. Carpenter, K.F.

Hirsch and J.R. Horner (Editors), Dinosaur Eggs and Babies. Cambridge University Press,

Cambridge, UK, pp. 151-183.

Vieira, L.G., Lima, F.C., Santos, A.L.Q., Mendonca, S.H.S.T., Moura, L.R., Lasbeck, J.R. and

Sebben, A., 2011. Description of embryonic stages in Melanosuchus niger (Spix, 1825)

(Crocodylia: Alligatoridae). Brazilian Journal of Morphological Science 28: 11-22.

Vila, B., Galobart, A., Oms, O., Poza, B. and Bravo, A.M., 2010. Assessing the nesting strategies

of Late Cretaceous titanosaurs: 3-D clutch geometry from a new megaloolithid eggsite.

Lethaia 43: 197-208.

Vila, B., Jackson, F.D., Fortuny, J., Selles, A.G. and Galobart, A., 2010. 3-D modelling of

megaloolithid clutches: insights about nest construction and dinosaur behaviour. PLoS One

5: 10.1371/journal.pone.0010362.

VillamarIn-Jurado, F. and Suarez, E., 2007. Nesting of the Black Caiman (Melanosuchus niger) in

Northeastern Ecuador. Journal of Herpetology 41: 164-167.

Visschedijk, A.H.J., Tazawa, H. and Piiper, J., 1985. Variability of shell conductance and gas

exchange of chicken eggs. Respiration Physiology 59: 339-346.

Vleck, C.M., Hoyt, D.F. and Vleck, D., 1979. Metabolism of avian embryos: patterns in altricial

and precocial birds. Physiological Zoology 52: 363-377. 437

Vleck, C.M., Vleck, D., Rahn, H. and Paganelli, C.V., 1983. Nest microclimate, water-vapor

conductance, and water-loss in heron and tern eggs. Auk 100: 76-83.

Vleck, D., Vleck, C.M. and Seymour, R.S., 1984. Energetics of embryonic development in the

megapode birds, Mallee Fowl Leipoa ocellata and Brush Turkey Alectura lathami.

Physiological Zoology 57: 444-456.

Wagner, K. and Seymour, R.S., 2001. Nesting climate and behaviour of Cape Barren geese

(Cereopsis novaehollandiae Latham). Australian Journal of Zoology 49: 155-170.

Waitkuwait, W.E., 1985. Investigations of the breeding biology of the West-african

slender-snouted crocodile Crocodylus cataphractus Cuvier, 1824. Amphibia-Reptilia 6:

387-399.

Walkinshaw, L.H., 1940. Summer life of the Sora Rail. Auk 57: 153-168.

Walkinshaw, L.H., 1950a. Incubation period of the Sandhill , Grus canadensis tabida. Auk

67: 513.

Walkinshaw, L.H., 1950b. The in the Bernard W. Baker Sanctuary, Michigan. Auk

67: 38-51.

Wang, D. and Zhou, S., 1995. The discovery of new typical dinosaur egg fossils in Xixia Basin.

Henan Geology 13: 262-267.

Wang, M., Zheng, X.T., O'Connor, J.K., Lloyd, G.T., Wang, X.L., Wang, Y., Zhang, X.M. and

Zhou, Z.H., 2015. The oldest record of Ornithuromorpha from the Early Cretaceous of

China. Nature Communications 6: 10.1038/Ncomms7987.

Wang, Q., Wang, X., Zhao, Z. and Jiang, Y., 2012. A new oofamily of dinosaur egg from the Upper

Cretaceous of Tiantai Basin, Zhejiang Province, and its mechanism of eggshell formation.

Chinese Science Bulletin 57: 3740-3747.

Wang, Q., Wang, X.L., Zhao, Z.K. and Jiang, Y.G., 2010. A new oogenus of Elongatoolithidae 438

from the Upper Cretaceous Chichengshan Formation of Tiantai Basin, Zhejiang Province.

Vertebrata Palasiatica 48: 111-118.

Wang, Q., Zan, S., Jin, L.Y. and Chen, J., 2006. A new oospecies, Dictyloolithus gongzhulingensis,

from the Early Cretaceous Quantou Formation in the central Jilin Province. Journal of Jilin

University (Earth Science Edition) 36: 153-157.

Wang, Q., Zhao, Z., Wang, X., Zhang, S.R. and Jiang, Y., 2013. New forms of dictyoolithids from

the Tiantai Basin, Zhejiang Province of China and a parataxonomic revision of the

dictyoolithids. Vertebrata PalAsiatica 51: 43-54.

Wang, Q., Zhao, Z., Wang, X.L. and Jiang, Y., 2011. New ootypes of dinosaur eggs from the Late

Cretaceous in Tiantai Basin, Zhejiang Province, China. Vertebrata Palasiatica 49: 446-449.

Wang, Q., Zhao, Z., Wang, X.L., Jiang, Y. and Zhang, S., 2010. A new oogenus of

macroelongatoolithid eggs from the Upper Cretaceous Chichengshan Formation of the

Tiantai Basin, Zhejiang Province and a revision of the macroelongatoolithids. Acta

Palaeontologica Sinica 49: 73-86.

Wang, Q., Zhao, Z.K., Wang, X.L., Li, N. and Zou, S.L., 2013. A new form of Elongatoolithidae,

Undulatoolithus pengi oogen. et oosp. nov. from Pingxiang, Jiangxi, China. Zootaxa 3746:

194-200.

Wang, S., Zhang, S., Sullivan, C. and Xu, X., 2016. Elongatoolithid eggs containing oviraptorid

(Theropoda, Oviraptorosauria) embryos from the Upper Cretaceous of Southern China.

BMC Evolutionary Biology 16: 1-21.

Warham, J., 1974. The breeding biology and behaviour of the snares crested penguin. Journal of

the Royal Society of New Zealand 4: 63-108.

Watabe, M., 2004. New dinosaur ovifauna from the Upper Cretaceous vertebrate fossil locality,

Abdrant Nuru, Central part of the Gobi desert, Mongolia. Hayashibara Museum of Natural 439

Sciences Research Bulletin 2: 15-27.

Watabe, M., Tsogtbaatar, K., Suzuki, S. and Saneyoshi, M., 2010. Geology of

dinosaur-fossil-bearing localities (Jurassic and Cretaceous: Mesozoic) in the Gobi Desert:

results of the HMNS-MPC Joint Paleontological Expedition. Hayashibara Museum of

Natural Sciences Research Bulletin 3: 41-118.

Watson, G.E., 1975. Birds of the Antarctic and Sub-Antarctic. Antarctic research series. American

Geophysical Union, Washington, 350 pp.

Weaver, R.L., 1943. Reproduction in English Sparrows. Auk 60: 62-74.

Webb, G.J.W., Beal, A.M., Manolis, S.C. and Dempsey, K.E., 1987. The effects of incubation

temperature on sex determination and embryonic development rate in Crocodylus

johnstoni and C. porosus. In: G.J.W. Webb, S.C. Manolis and P.J. Whitehead (Editors),

Wildlife Management: Crocodiles and Alligators. Surrey Beatty and Sons, Chipping

Norton, Australia, pp. 507-531.

Webb, G.J.W., Buckworth, R. and Manolis, S.C., 1983. Crocodylus johnstoni in the Mckinlay

River N.T. VI. Nesting biology. Australian Wildlife Research 10: 607-637.

Webb, G.J.W. and Cooper-Preston, H., 1989. Effects of incubation temperature on crocodiles and

the evolution of reptilian . American Zoologist 29: 953-971.

Webb, G.J.W., Messel, H. and Magnusson, W., 1977. Nesting of Crocodylus porosus in Arnhem

Land, Northern Australia. Copeia: 238-249.

Webb, G.J.W., Sack, G.C., Buckworth, R. and Manolis, S.C., 1983. An examination of Crocodylus

porosus nests in two northern Australian fresh water swamps, with an analysis of embryo

mortality. Australian Wildlife Research 10: 571-605.

Weir, D.G., 1973. Status and habits of Megapodius pritchardii. Wilson Bulletin 85: 79-82.

Weishampel, D.B., Barrett, P.M., Coria, R.A., Loeuff, J., Xu, X., Xijin, Z., Sahni, A., Gomani, 440

E.M.P. and Noto, C.R., 2004. Dinosaur distribution. In: D.B. Weishampel, P. Dodson and H.

Osmolska (Editors), Dinosauria, 2nd Ed. University of California Press, Barkeley and Los

Angels, pp. 517-606.

Weishampel, D.B., Fastovsky, D.E., Watabe, M., Varricchio, D., Jackson, F., Tsogtbaatar, K. and

Barsbold, R., 2008. New oviraptorid embryos from Bugin-Tsav,

(Upper Cretaceous), Mongolia, with insights into their habitat and growth. Journal of

Vertebrate Paleontology 28: 1110-1119.

Welsh, E. and Sankey, J.T., 2005. First dinosaur eggshells from Texas, USA: Aguja Formation

(Late Campanian), Big Bend National Park. In: J.T. Sankey and S. Baszio (Editors),

Vertebrate Microfossil Assembrages: Their Role in Paleoecology and Paleobiogeography.

Indiana University Press, Bloomington and Indianapolis, pp. 166-177.

Werner, J. and Griebeler, E.M., 2013. New insights into non-avian dinosaur reproduction and their

evolutionary and ecological implications: linking fossil evidence to allometries of extant

close relatives. PLoS One 8: 10.1371/journal.pone.0072862.

Whistler, H.F.Z.S., 1949. Popular Handbook of Indian Birds. Oliver and Boyd, Edimburgh and

London, UK, 414 pp.

Whitaker, R. and Whitaker, Z., 1976. Collection and hatching of Marsh Crocodile (C. palustris)

eggs. Journal of Bombay Natural History Society 73: 403-407.

White, D.W. and Kennedy, E.D., 1997. Effect of egg covering and habitat on nest destruction by

House Wrens. Condor 99: 873-879.

Whitehead, P.J. and Seymour, R.S., 1990. Patterns of metabolic rate in embryonic crocodilians

Crocodylus johnstoni and Crocodylus porosus. Physiological Zoology 63: 334-352.

Whittow, G.C., 1980. Physiological and ecological correlates of prolonged incubation in sea birds.

American Zoologist 20: 427-436. 441

Whittow, G.C., 1983. Physiological ecology of incubation in tropical seabirds. Studies in Avian

Biology 8: 47-72.

Whittow, G.C., 1997. Wedge-tailed Shearwater (Puffinus pacificus). In: A. Poole and F. Gill

(Editors), The Birds of North America, No. 305. The Academy of Natural Sciences,

Philadelphia, and the American Ornithologists' Union, Washington, D.C., pp. 1-24.

Whittow, G.C., 2001. Incubation weight loss of Christmas Shearwater eggs on Christmass Island,

Pacific Ocean. Notornis 48: 175-176.

Whittow, G.C., Ackerman, R.A., Paganelli, C.V. and Pettit, T.N., 1982. Pre-pipping water loss

from the eggs of the Wedge-tailed Shearwater. Comparative Biochemistry and Physiology

A 72: 29-34.

Whittow, G.C., Grant, G.S. and Flint, E.N., 1985. Egg water loss, shell water-vapor conductance,

and the incubation period of the Gray-backed Tern (Sterna lunata). Condor 87: 269-272.

Whittow, G.C., Grant, G.S. and Pettit, T.N., 2003. Water loss from eggs of the Great .

Wilson Bulletin 115: 99-101.

Whittow, G.C. and Pettit, T.N., 2000. Egg dimensions and shell characteristics of Bulwer's Petrels,

Bulweria bulwerii, on Laysan Island, northwestern Hawaiian islands. Pacific Science 54:

183-188.

Whittow, G.C., Pettit, T.N., Ackerman, R.A. and Paganelli, C.V., 1989. The regulation of

water-loss from the eggs of the Red-footed Booby (Sula sula). Comparative Biochemistry

and Physiology A 93: 807-810.

Whittow, G.C., Simons, T.R. and Pettit, T.N., 1984. Water loss from the eggs of a tropical sea bird

(Pterodroma phaeopygia) at high altitude. Comparative Biochemistry and Physiology A

78: 537-540.

Wiemann, J., Yang, T., Sander, P.N.N., Schneider, M., Engeser, M., Kath-Schorr, S., Müller, C.E. 442

and Sander, P.M., 2015. The blue-green eggs of dinosaurs: how fossil metabolites provide

insights into the evolution of bird reproduction. PeerJ PrePrints 3: e1323

https://dx.doi.org/10.7287/peerj.preprints.1080v1.

Wiles, G.J. and Conry, P.J., 2001. Characteristics of nest mounds of Micronesian Megapodes in

Palau. Journal of Field Ornithology 72: 267-275.

Wiley, R.H. and Lee, D.S., 1998. Long-tailed Jaeger (Stercorarius longicaudus). In: A. Poole and

F. Gill (Editors), The Birds of North America, No. 365. The Birds of North America, Inc.,

Philadelphia, pp. 1-24.

Wilken, D.W., Staton, M.A., Vernon, B.P. and Bon, B., 1992. Artificial incubation of saltwater

crocodile (Crocodylus porosus) eggs. In: Proceedings of the 11th Working Meeting of the

Crocodile Specialist Group of the Species Survival Comission of the IUCN - the World

Conservation Union. IUCN, Gland, Switzerland, pp. 276.

Williams, D.L.G., Seymour, R.S. and Kerourio, P., 1984. Structure of fossil dinosaur eggshell from

the Aix Basin, France. Palaeogeography Palaeoclimatology Palaeoecology 45: 23-37.

Williams, M., 1979. The social structure, breeding and population dynamics of Paraside Shelduck

in the Gisborne-East Coast District. Notornis 26: 213-272.

Williams, T.D., 1995. The Penguins: Spheniscidae. Oxford University Press, Oxford, UK, 295 pp.

Wilson, J.A., Mohabey, D.M., Peters, S.E. and Head, J.J., 2010. Predation upon hatchling

dinosaurs by a new snake from the Late Cretaceous of India. PLoS Biology 8:

10.1371/journal.pbio.1000322

Wink, C.S. and Elsey, R.M., 1994. Morphology of shells from viable and nonviable eggs of the

Chinese Alligator (Alligator sinensis). Journal of Morphology 222: 103-110.

Wink, C.S., Elsey, R.M. and Bouvier, M., 1990. Porosity of eggshells from wild and captive,

Pen-reared Alligators (Alligator mississippiensis). Journal of Morphology 203: 35-39. 443

Wink, M. and Heidrich, P., 2000. Molecular systematics of owls (Strigiformes) based on

DNA-sequences of the mitochondrial cytochrome b gene. In: R.D. Chancellor and B.-U.

Meyburg (Editors), Raptors at Risk. WWGBP, Hancock House, London, UK, pp. 819-828.

Wolf, B.O., 1997. Black Phoebe (Sayornis nigricans). In: A. Poole and F. Gill (Editors), The Birds

of North America, No. 268. The Academy of Natural Sciences, Philadelphia, and the

American Ornithologists' Union, Washington, D.C., pp. 1-20.

Woodall, P.F. and Parry, D.F., 1982. Water-loss during incubation in Red Bishop (Euplectes orix)

eggs. South African Journal of Zoology 17: 75-78.

Woolfenden, G.E., 1978. Growth and survival of Young Florida Scrub Jays. Wilson Bulletin 90:

1-18.

Woolfenden, G.E. and Fitzpatrick, J.W., 1996. Florida Scrub-Jay (Aphelocoma coerulescens). In:

A. Poole and F. Gill (Editors), The Birds of North America, No. 228. The Academy of

Natural Sciences, Philadelphia, and the American Ornithologists' Union, Washington, D.C.,

pp. 1-28.

Wooller, R.D. and Dunlop, J.N., 1980. The use of simple measurements to determine the age of

Silver Gull eggs. Australian Wildlife Research 7: 113-115.

Wright, T.F., Schirtzinger, E.E., Matsumoto, T., Eberhard, J.R., Graves, G.R., Sanchez, J.J.,

Capelli, S., Mueller, H., Scharpegge, J., Chambers, G.K. and Fleischer, R.C., 2008. A

multilocus molecular phylogeny of the parrots (Psittaciformes): support for a Gondwanan

origin during the Cretaceous. Molecular Biology and Evolution 25: 2141-2156.

Wu, J. and Peng, A., 2003. Identification, paleogeography, and paleoclimate analyses of fossil

dinosaur eggs from Baiyuan-Zhenyuan, Pingxiang, Jiangxi Province. Relics From South 3:

42-44.

Xie, J.-F., Zhang, S.-K., Jin, X.-S., Li, D.-Q. and Zhou, L.-Q., 2016. A new type of dinosaur eggs 444

from Early Cretaceous of Gansu Province, China. Vertebrata PalAsiatica 54: 1-10.

Xu, L., 2007. The characters and protection of dinosaurs from Henan Province. Ph.D. Thesis,

China University of Geosciences, Beijing, China, 120 pp.

Xu, X., Zhou, Z., Dudley, R., Mackem, S., Chuong, C.-M., Erickson, G.M. and Varricchio, D.J.,

2014. An integrative approach to understanding bird origins. Science 346:

10.1126/science.1253293.

Yamashina, Y., 1961. Birds in Japan: a Field Guide. Tokyo News Service, Tokyo, Japan, 233 pp.

Yang, R., Wu, X.B., Yan, P., Su, X. and Yang, B.H., 2010. Complete mitochondrial genome of Otis

tarda (Gruiformes: Otididae) and phylogeny of Gruiformes inferred from mitochondrial

DNA sequences. Molecular Biology Reports 37: 3057-3066.

Young, C.C., 1954. Fossil reptilian eggs from Laiyang, Shantung, China. Acta Palaeontologica

Sinica 34: 371-388.

Young, C.C., 1965. Fossil eggs from Nanhsiung, Kwangtung and Kanchou, Kiangsi. Vertebrata

Palasiatica 9: 141-170.

Yun, C.-S. and Yang, S.-Y., 1997. Dinosaur eggshells from the Hasandong Formation,

Gyeongsang Supergroup, Korea. Journal of the Paleontological Society of Korea 13:

21-36.

Zar, J.H., 1968. Calculation and miscalculation of the allometric equation as a model in biological

data. Bioscience 18: 1118–1120.

Zelenitsky, D.K., 2004. Description and phylogenetic implications of extant and fossil oologic

remains. Ph.D. Thesis, University of Calgary, Calgary, Canada, 364 pp.

Zelenitsky, D.K., Carpenter, K. and Currie, P.J., 2000. First record of elongatoolithid theropod

eggshell from North America: the Asian oogenus Macroelongatoolithus from the Lower

Cretaceous of Utah. Journal of Vertebrate Paleontology 20: 130-138. 445

Zelenitsky, D.K. and Hills, L.V., 1996. An egg clutch of Prismatoolithus levis oosp. nov. from the

Oldman Formation (Upper Cretaceous), Devil's Coulee, southern Alberta. Canadian

Journal of Earth Sciences 33: 1127-1131.

Zelenitsky, D.K. and Hills, L.V., 1997. Normal and pathological eggshells of Spheroolithus

albertensis, oosp. nov., from the (Judith River Group, late Campanian),

southern Alberta. Journal of Vertebrate Paleontology 17: 167-171.

Zelenitsky, D.K., Hills, L.V. and Currie, P.J., 1996. Parataxonomic classification of ornithoid

eggshell fragments from the Oldman Formation (Judith River Group; Upper Cretaceous),

southern Alberta. Canadian Journal of Earth Sciences 33: 1655-1667.

Zelenitsky, D.K. and Modesto, S.P., 2002. Re-evaluation of the eggshell structure of eggs

containing dinosaur embryos from the Lower Jurassic of South Africa. South African

Journal of Science 98: 407-408.

Zelenitsky, D.K. and Sloboda, W., 2005. Eggshells. In: P.J. Currie and E.B. Koppelhus (Editors),

Dinosaur Provincial Park: a Spectacular Ancient Ecosystem Revealed. Indiana University

Press, Bloomington, pp. 398-404.

Zelenitsky, D.K. and Therrien, F., 2008a. Phylogenetic analysis of reproductive traits of

maniraptoran theropods and its implications for egg parataxonomy. Palaeontology 51:

807-816.

Zelenitsky, D.K. and Therrien, F., 2008b. Unique maniraptoran egg clutch from the Upper

Cretaceous Two Medicine Formation of Montana reveals theropod nesting behaviour.

Palaeontology 51: 1253-1259.

Zeng, D. and Zhang, J., 1979. On the dinosaurian eggs from the western Dongting Basin, Hunan.

Vertebrata Palasiatica 17: 131-136.

Zhang, S.-K., 2010. A parataxonomic revision of the Cretaceous faveoloolithid eggs of China. 446

Vertebrata PalAsiatica 48: 203-219.

Zhang, W. and Huang, P., 1999. The strata with fossil dinosaur eggs and their distribution in

Guangdong. Journal of Stratigraphy 23: 89-95.

Zhao, H. and Zhao, Z., 1998. Dinosaur eggs from Xichuan Basin, Henan Province. Vertebrata

Palasiatica 36: 282-296.

Zhao, H. and Zhao, Z., 1999. A new form of elongatoolithid dinosaur eggs from the Lower

Cretaceous Shahai Formation of Heishan, Liaoning Province. Vertebrata Palasiatica 37:

278-284.

Zhao, Z., 1979a. Discovery of the dinosaurian eggs and footprint from Neixiang County, Henan

Province. Vertebrata PalAsiatica 17: 304-309.

Zhao, Z., 1979b. Progress in the research of dinosaur eggs. In: Institute of Vertebrate Paleontology,

Paleoanthropology and Nanjing Institute of Paleontology (Editor), Mesozoic and Cenozoic

Red Beds of South China Selected Papers from the "Cretaceous - Tertiary Workshop,

Nanxiong, Guangdong Province". Science Press, Beijing, China, pp. 330-340.

Zhao, Z., 1994. Dinosaur eggs in China; on the structure and evolution of eggshells. In: K.

Carpenter, K.F. Hirsch and J.R. Horner (Editors), Dinosaur Eggs and Babies. Cambridge

University Press, Cambridge, UK, pp. 184-203.

Zhao, Z. and Ding, S., 1976. Discovery of the dinosaurian eggshells from Alxa, Ningxia and its

stratigraphic significance. Vertebrata PalAsiatica 14: 42-44.

Zhao, Z. and Jiang, Y., 1974. Microscopic studies on the dinosaurian egg-shells from Laiyang,

Shantung Province. Scientia Sinica 17: 73-83.

Zhao, Z. and Li, R., 1988. A new structural type of the dinosaur eggs from Anlu Country, Hubei

Province. Vertebrata PalAsiatica 26: 107-115.

Zhao, Z. and Li, R., 1993. First record of Late Cretaceous hypsilophodontid eggs from Bayan 447

Manduhu, Inner Mongolia. Vertebrata Palasiatica 31: 77-84.

Zhao, Z., Ye, J., Li, H., Zhao, Z. and Yan, Z., 1991. Extinction of the dinosaurs across the

Cretaceous - Tertiary boundary in Nanxiong Basin, Guangdong Province. Vertebrata

Palasiatica 29: 1-20.

Zhao, Z., Zhang, S., Wang, Q. and Wang, X.L., 2013. Dinosaur diversity during the transition

between the middle and late parts of the Late Cretaceous in eastern Shangdong Province,

China: evidence from dinosaur eggshells. Chinese Science Bulletin 58: 4663-4669.

Zhao, Z.K., 1975. Microstructures of dinosaurian eggshells of Nanxiong, Guangdong, and the

problem in egg classification. Vertebrata Palasiatica 13: 105-117.

Zhao, Z.K., Mao, X.Y., Chai, Z.F., Yang, G.C., Kong, P., Ebihara, M. and Zhao, Z.H., 2002. A

possible causal relationship between extinction of dinosaurs and K/T enrichment in

the Nanxiong Basin, South China: evidence from dinosaur eggshells. Palaeogeography

Palaeoclimatology Palaeoecology 178: 1-17.

Zheng, X., O'Connor, J., Huchzermeyer, F., Wang, X., Wang, Y., Wang, M. and Zhou, Z., 2013.

Preservation of ovarian follicles reveals early evolution of avian reproductive behaviour.

Nature 495: 507-511.

Zhou, X., Ren, Y., Xu, S. and Guan, K., 1998. Dinosaur eggs of the Late Cretaceous from the

Qinglongshan region, Yunxian County, Hubei Province. Hubei Geology and Mineral

Resources 12: 1-8.

Zimmermann, K. and Hipfner, J.M., 2007. Egg size, eggshell porosity, and incubation period in the

marine bird family. Auk 124: 307-315.

Zou, S.L., Wang, Q. and Wang, X.L., 2013. A new oospecies of parafaveoloolithids from the

Pingxiang Basin, Jiangxi Province of China. Vertebrata PalAsiatica 51: 102-106.

448

NRC Research Press

Subscriber access provided by University of Login Register Shibboleth Mobile Cart Calgary

Advanced Search A division of Canadian Science Publishing a not-for-profit publisher All Journals

Home CSP Journals Books Compilations Open Access Authors Librarians Societies About the Press Contact Français

Home > Authors' rights

Services Authors' rights As of 2009, copyright of all articles in NRC Research Press journals remains with the authors. Copyright of all Submissions articles published prior to 2009 is held by Canadian Science Publishing (operating as NRC Research Press) or its licensors (also see our Copyright and Reuse of Content information). Just-IN Under the terms of the license to publish granted to NRC Research Press, authors retain the following Reprints & permissions rights:

Video abstracts 1. To post a copy of their submitted manuscript (pre-print) on their own website, an institutional repository, a preprint server, or their funding body's designated archive (no embargo period). Plain language Publication on a preprint server prior to submitting to an NRC Research Press journal does not summaries constitute “prior publication”. 2. To post a copy of their accepted manuscript (post-print) on their own website, an institutional Enrich your article repository, a preprint server, or their funding body's designated archive (no embargo period). Authors who archive or self-archive accepted articles are asked to provide a hyperlink from the manuscript to How to Deposit to the Journal's website. Tspace or Another Institution Repository 3. Authors, and any academic institution where they work at the time, may reproduce their manuscript for the purpose of course teaching. Publishing Toolkit 4. Authors may reuse all or part of their manuscript in other works created by them for non-commercial for authors purposes, provided the original publication in an NRC Research Press journal is acknowledged through a note or citation. License to publish forms (copyright agreements for publishing) These authors’ rights ensure that NRC Research Press journals are compliant with open access policies of research funding agencies, including the Canadian Institutes of Health Research, the Natural Sciences and Engineering Research Council of Canada, the US National Institutes of Health, the Wellcome Trust, the UK Permission forms Medical Research Council, l'Institut national de la santé et de la recherche médicale in France, and others. Your responsibilities as In support of authors who wish or need to sponsor open access to their published research articles, NRC an author (Publishing Research Press also offers a Open Access (OpenArticle) option. Policy) The above rights do not extend to copying or reproduction of the full article for commercial purposes. Files and graphics Authorization to do so may be obtained by clicking on the "Reprints & Permissions" link in the Article Tools menu of the article in question or under license by Access Copyright. The Article Tools menu is accessible Authors' rights through the full-text article or abstract page. General ethical guidelines

English-improvement services

Successful publishing strategies for authors

What we're blogging about

Open Access: Reaching Your Audience Posted: Oct 25, 2016

By Michael Donaldson In honour of Open Access Week, we explore how open access removes barriers to scholarly research which…

Celebrating Open Access Week 2016 Posted: Oct 24, 2016

http://www.nrcresearchpress.com/page/authors/information/rights[11/2/2016 1:31:23 PM] PLOS ONE: accelerating the publication of peer-reviewed science

plos.orgcreate accountsign in

PublishAboutBrowse 

advanced search

Licenses and Copyright The following policy applies to all PLOS journals, unless otherwise noted. What Can Others Do with My Original Article Content? PLOS applies the Creative Commons Attribution (CC BY) license to articles and other works we publish. If you submit your paper for publication by PLOS, you agree to have the CC BY license applied to your work. Under this Open Access license, you as the author agree that anyone can reuse your article in whole or part for any purpose, for free, even for commercial purposes. Anyone may copy, distribute, or reuse the content as long as the author and original source are properly cited. This facilitates freedom in re-use and also ensures that PLOS content can be mined without barriers for the needs of research. May I Use Content Owned by Someone Else in My Article? If you have written permission to do so, yes. If your manuscript contains content such as photos, images, figures, tables, audio files, videos, etc., that you or your co-authors do not own, we will require you to provide us with proof that the owner of that content (a) has given you written permission to use it, and (b) has approved of the CC BY license being applied to their content. We provide a form you can use to ask for and obtain permission from the owner. Download the form (PDF). If you do not have owner permission, we will ask you to remove that content and/or replace it with other content that you own or have such permission to use. Don't assume that you can use any content you find on the Internet, or that the content is fair game just because it isn't clear who the owner is or what license applies. It's up to you to ascertain what rights you have —if any—to use that content. May I Use Article Content I Previously Published in Another Journal? Many authors assume that if they previously published a paper through another publisher, they own the rights to that content and they can freely use that content in their PLOS paper, but that’s not necessarily the case – it depends on the license that covers the other paper. Some publishers allow free and unrestricted re- use of article content they own, such as under the CC BY license. Other publishers use licenses that allow re-use only if the same license is applied by the person or publisher re-using the content. If the paper was published under a CC BY license or another license that allows free and unrestricted use, you may use the content in your PLOS paper provided that you give proper attribution, as explained above. If the content was published under a more restrictive license, you must ascertain what rights you have under that license. At a minimum, review the license to make sure you can use the content. Contact that publisher if you have any questions about the license terms – PLOS staff cannot give you legal advice about your rights to use third-party content. If the license does not permit you to use the content in a paper that will be covered by an unrestricted license, you must obtain written permission from the publisher to use the content in your PLOS paper. Please do not include any content in your PLOS paper which you do not have rights to use, and always give proper attribution. What Are Acceptable Licenses for Data Repositories? If any relevant accompanying data is submitted to repositories with stated licensing policies, the policies should not be more restrictive than CC BY. Removal of Content Used Without Clear Rights PLOS reserves the right to remove any photos, captures, images, figures, tables, illustrations, audio and http://journals.plos.org/plosone/s/licenses-and-copyright[11/2/2016 1:38:34 PM] PLOS ONE: accelerating the publication of peer-reviewed science

video files, and the like, from any paper, whether before or after publication, if we have reason to believe that the content was included in your paper without permission from the owner of the content. How Does One Give Proper Attribution for Use of Content? When citing a PLOS research article, use the “Vancouver style”, as outlined in our Submission Guidelines. For example: Kaltenbach LS et al. (2007) Huntingtin Interacting Are Genetic Modifiers of Neurodegeneration. PLOS Genet 3(5): e82. doi:10.1371/journal.pgen.0030082. When citing non-article content from a PLOS website (e.g., blog content), provide a link to the content, and cite the title and author(s) of that content. For examples of proper attribution to other types of content, see websites such as Open.Michigan.

Publications PLOS Biology PLOS Medicine Privacy PolicyTerms of UseAdvertiseMedia InquiriesPLOS Computational Biology PLOS Currents PLOS Genetics PLOS Pathogens PLOS ONE PLOS Neglected Tropical Diseases

plos.org Blogs Collections Send us feedback Help using this site LOCKSS California (US) corporation #C2354500, based in San Francisco

http://journals.plos.org/plosone/s/licenses-and-copyright[11/2/2016 1:38:34 PM] Creative Commons — Attribution 4.0 International — CC BY 4.0

Attribution 4.0 International (CC BY 4.0)

This is a human-readable summary of (and not a substitute for) the license.

Disclaimer

You are free to:

Share — copy and redistribute the material in any medium or format

Adapt — remix, transform, and build upon the material

for any purpose, even commercially.

The licensor cannot revoke these freedoms as long as you follow the license terms.

Under the following terms:

Attribution — You must give appropriate credit, provide a link to the license, and indicate if changes were made. You may do so in any reasonable manner, but not in any way that suggests the licensor endorses you or your use.

No additional restrictions — You may not apply legal terms or technological measures that legally restrict others from doing anything the license permits.

Notices:

You do not have to comply with the license for elements of the material in the public domain or where your use is permitted by an applicable exception or limitation.

No warranties are given. The license may not give you all of the permissions necessary for your intended use. For example, other rights such as publicity, privacy, or moral rights may limit how you use the material.

Learn more about CC licensing, or use the license for your own material.

This page is available in the following languages: Castellano Castellano (España) Català Dansk Deutsch English Esperanto français Galego hrvatski Indonesia Italiano Latviski Lietuvių Magyar Melayu Nederlands Norsk polski Português Português (BR) Português (Portugal) română Suomeksi svenska Türkçe íslenska česky Ελληνικά Беларуская русский українська 中文 日本語 華語 (台灣) 한국어

https://creativecommons.org/licenses/by/4.0/[11/2/2016 1:38:19 PM]