City University of New York (CUNY) CUNY Academic Works

All Dissertations, Theses, and Capstone Projects Dissertations, Theses, and Capstone Projects

9-2017

Climate-Driven Habitat Shifts in South America: Biogeography, Historical Demography, and Population Genomics of Anole

Ivan Prates The Graduate Center, City University of New York

How does access to this work benefit ou?y Let us know!

More information about this work at: https://academicworks.cuny.edu/gc_etds/2414 Discover additional works at: https://academicworks.cuny.edu

This work is made publicly available by the City University of New York (CUNY). Contact: [email protected]

CLIMATE-DRIVEN HABITAT SHIFTS IN SOUTH AMERICA:

BIOGEOGRAPHY, HISTORICAL DEMOGRAPHY, AND POPULATION GENOMICS OF ANOLE LIZARDS

by

Ivan Prates

A dissertation submitted to the Graduate Faculty in Biology in partial fulfillment of the

requirements for the degree of Doctor of Philosophy, The City University of New York

2017

© 2017

IVAN PRATES

All Rights Reserved

ii

CLIMATE-DRIVEN HABITAT SHIFTS IN SOUTH AMERICA:

BIOGEOGRAPHY, HISTORICAL DEMOGRAPHY, AND POPULATION GENOMICS OF ANOLE LIZARDS

by

Ivan Prates

This manuscript has been read and accepted for the Graduate Faculty in Biology in satisfaction

of the dissertation requirement for the degree of Doctor of Philosophy.

______Date Chair of Examining Committee Dr. Ana Carolina Carnaval, City College of New York

______Date Executive Officer Dr. Cathy Savage-Dunn

______Dr. Michael J. Hickerson, City College of New York

______Dr. Frank T. Burbrink, American Museum of Natural History

______Dr. Christopher J. Raxworthy, American Museum of Natural History

______Dr. Jason Munshi-South, Fordham University

Supervising Committee

The City University of New York

iii

Abstract

CLIMATE-DRIVEN HABITAT SHIFTS IN SOUTH AMERICA:

BIOGEOGRAPHY, HISTORICAL DEMOGRAPHY, AND POPULATION GENOMICS OF ANOLE LIZARDS

by

Ivan Prates

Advisor: Ana Carolina Carnaval, Ph.D.

Shifts in the geographic distribution of habitats over time can promote dispersal and vicariance, thereby influencing large-scale biogeographic patterns and ecological processes. The establishment of corridors of suitable habitat across previously disjunct yet ecologically similar regions has been widely linked to climate change over time. Such climate-mediated habitat changes may have played a key role in the assembly of tropical biotas, including those of

Amazonia and the coastal Atlantic Forest in South America, two of the most diverse ecosystems on Earth. The history of biogeographic associations between tropical rainforest habitats in response to former climatic shifts, as well as their consequences for the distribution and genetic composition of forest taxa, is the theme of this doctoral dissertation. To study the historical biogeography of these forest blocks, the four chapters presented here collect and analyze genetic data from South American of true anoles (Anolis, ) and, to a lesser extent, bush anoles (, ).

To improve inferences about historical biogeography and the drivers of diversification in

South America, my dissertation begins with a revisionary phylogenetic study of poorly-known rainforest anole species. Chapter 1 provides an updated molecular phylogeny of a highly diverse clade within the genus Anolis known as Dactyloa. My phylogenetic analyses find A. phyllorhinus to be nested within the punctatus group, suggesting independent evolution of a rostral proboscis

iv in Anolis. Moreover, A. philopunctatus is recovered as nested within A. punctatus, with limited genetic divergence between distinct dewlap phenotypes. The most recent common ancestor of the Dactyloa and Norops clades of Anolis dates back to the Eocene. Most Amazonian species within these two clades diverged in the Miocene, while Amazonian species of Polychrus diverged in the Pliocene. Importantly, the temporal framework provided is this study yielded much-needed estimates of divergence times and mutation rates for my subsequent phylogeographic and population genetic analyses of historical biogeographic connections between Neotropical rainforests based on anoles.

Focusing on the southern montane region of the Atlantic Rainforest, where former biogeographic connections with other South American regions have been invoked to explain the affinities of a number of endemic taxa, Chapter 2 investigates the phylogenetic relationships of

Anolis nasofrontalis and A. pseudotigrinus, known from few specimens collected more than 40 years ago. A comprehensive phylogenetic analysis recovers six main clades within the Dactyloa clade of Anolis, five of which were previously referred to as species series (aequatorialis, heterodermus, latifrons, punctatus, roquet). A sixth clade clustered A. nasofrontalis and A. pseudotigrinus with A. dissimilis from western Amazonia, A. calimae from the Andes, A. neblininus from the Guiana Shield, and two undescribed Andean taxa. I therefore define a sixth species series within Dactyloa: the neblininus series. Close phylogenetic relationships between highly disjunct, narrowly-distributed anoles suggest that patches of suitable habitat connected the southern Atlantic Forest to western South America during the Miocene, in agreement with the age of former connections between the central Andes and the Brazilian Shield as a result of

Andean orogeny. The data also support the view of recurrent evolution (or loss) of a twig anole- like phenotype in mainland anoles, in apparent association with the occurrence in montane

v settings.

In Chapter 3, I employ sequence data at multiple loci from three co-distributed arboreal lizards (Anolis punctatus, A. ortonii, and Polychrus marmoratus) to infer historical relationships among disjunct Amazonian and Atlantic Forest populations and to test alternative historical demographic scenarios of colonization and vicariance using coalescent simulations and

Approximate Bayesian Computation (ABC). Data from the better-sampled Anolis species support colonization of the Atlantic Forest from eastern Amazonia, while hierarchical ABC suggests that the three species colonized the Atlantic Forest synchronously during the mid-

Pleistocene. I find support of population bottlenecks associated to founder events in the two

Anolis, but not in P. marmoratus, consistently with their distinct ecological tolerances. These findings support that climatic fluctuations provided key opportunities for dispersal and forest colonization in eastern South America through the cessation of environmental barriers. Evidence of species-specific histories strengthens assertions that biological attributes play a role in responses to shared environmental change.

In Chapter 4, I assess whether range expansions across ecological gradients in South

America are associated with the selection of genotypes that confer physiological adaptation. For that, I carry a comparative population genomic study based on restriction enzyme associated

DNA markers from the anoles Anolis ortonii and A. punctatus, which colonized the Atlantic

Rainforest from Amazonia. Clustering analyses indicate strong genetic differentiation between

Atlantic Forest and Amazonian samples in both anole species, while historical demographic inference recovers large effective population sizes and Pleistocene divergences between populations in the two forest blocks. Low levels of gene flow indicate very limited post- divergence exchange of migrants between Amazonian and Atlantic forests, in spite of the

vi presumed opportunities of connectivity provided by cyclic periods of increased humidity and forest expansions. Genome-environment association analyses recover eight protein-coding loci in

A. ortonii and 14 in A. punctatus whose frequency is significantly correlated to environmental gradients, consistent with a scenario of local adaptation. These loci are involved with metabolic and regulation processes, performing molecule binding, transcription, hydrolase, and transporter activities. My results suggest a role of functional genomic differentiation and local adaptation associated with range expansions across heterogeneous landscapes in widely distributed species.

The establishment of allele frequency differences in response to climatic regimes may have been favored by large effective population sizes and low migration rates across regions, consistently with proposed interactions between natural selection and the demographic trajectory of populations during the process of adaptation.

vii

Acknowledgements

Thanks so much to my dissertation supervisor Ana C. Carnaval for being an incredibly supportive and inspiring mentor over the last six years. Ana has been extremely generous with her time, providing advice for project development, patiently teaching me scientific writing, and showing interest in my professional trajectory. From day one, Ana has treated me as a colleague and stimulated independent thinking. As an advisor, she has been fair and recognized hard work.

Ana has contributed key resources for me to develop research that is aligned with my scientific interests and career goals. She has used her academic position to help me integrate with the scientific community and to assist me navigate institutional bureaucracy. With Ana's support, my

PhD was a fruitful and gratifying experience, and I hope to continue to have her as a mentor for many years to come.

Thanks to my dissertation committee members Michael J. Hickerson, Frank T. Burbrink,

Jason Munshi-South, Christopher J. Raxworthy, and previous member Catherine H. Graham, as well as to Brian T. Smith and Robert Anderson, for providing numerous constructive comments, methodological insights, and a rich environment for the exchange of ideas that greatly helped shaping my views as a scientist.

Thanks to Carnaval Lab members Danielle Rivera, Maria L. Strangas, Andrea Paz, Jason

Brown, Mariana Vasconcellos, Leyla Hernandez, Barbara Rizzo, and Zoe Spanos, who provided a fun and scientifically rich workspace, trained me in molecular methods, taught me actual

English, and provided crucial comments on all my drafts.

Thanks to my fellow PhD students and postdocs at CUNY and the AMNH Alexander

Xue, Edward A. Myers, Gregory Thom, Isaac Overcast, Carolina Machado, Marcelo Gehara,

Diego Alvarado Serrano, and Marc Tollis for methodological insights and assistance with the

viii implementation of complicated analyses.

Muito obrigado to my Brazilian collaborators and friends Paulo R. Melo Sampaio, Mauro

Teixeira Jr., Francisco Dal Vecchio, Sérgio M. de Souza, Renato Recoder, Roberta Damasceno,

Leandro Drummond, Marco Sena, and especially my de facto co-advisor Miguel T. Rodrigues.

Beyond providing samples and fieldwork support, they taught me a great deal about and diversity during the formidable 12 months that we spent doing biological inventories over the course of my PhD.

Thanks to my indispensable New York friends Danielle Rivera, Pedro V. Peloso, Silvia

Pavan, José M. Padial, Beatriz Corradini, and parceiro Gustavo M. Silva for all the beers, bobós, and dance competitions, as well as discussions about philosophy of science, diversity in science and society, career, and . Thanks to my roommates Peter Galante, Nicolle A. Konai, and

Leonardo Duran, who helped me cope with living far from home. Thanks so much to my exceptional brother Victor Prates for taking care of our mother and countless annoying matters while I was pursuing a PhD overseas.

Thanks so much to my extraordinary partner Anna Penna for transforming the terrifying prospect of a short-lived long-distance relationship into a beautiful history of support, exchange, and immeasurable fun across 13 countries as we grew closer over the past six years. Thanks also for patiently teaching me all that I know about R scripting.

Key samples and specimen information used in this dissertation research were provided by Alberto Carvalho and Hussan Zaher (Museu de Zoologia da Universidade de São Paulo),

Alexander Haas and Jakob Hallermann (Zoologisches Museum Hamburg), Ana Prudente and

Annelise D’Angiolella (Museu Paraense Emílio Goeldi), Georg Gassner (Naturhistorisches

Museum Wien), Helio Fernandes and Rose Kollmann (Museu de Biologia Professor Mello

ix

Leitão), and Frederick Sheldon (Louisiana State University Museum of Natural Science).

Fieldwork in was facilitated by Moisés Barbosa (Fazenda Experimental Catuaba,

Universidade Federal do ), Carlos Rangel (Parque Nacional Pacaás Novos), Cleide Rezende

(Parque Nacional da Serra do Divisor), Thomaz Novaes and Angelica Tomás (Reserva Biológica

Augusto Ruschi), and Museu de Biologia Professor Mello Leitão (Instituto Nacional da Mata

Atlântica). Collection permits were issued by Instituto Chico Mendes de Conservação da

Biodiversidade.

This research was funded by the National Science Foundation, CUNY Graduate Center,

CUNY Research Foundation, Fundação de Amparo à Pesquisa do Estado de São Paulo, and

Conselho Nacional de Desenvolvimento Científico e Tecnológico.

Chapters one to three of this dissertation have been published in scientific journals, as follows:

Chapter 1: Prates I., Rodrigues M.T., Melo-Sampaio P.R., Carnaval A.C. (2015). Phylogenetic relationships of Amazonian anole lizards (Dactyloa): taxonomic implications, new insights about phenotypic evolution and the timing of diversification. Molecular Phylogenetics and Evolution,

82: 258-268.

Chapter 2: Prates I., Melo-Sampaio P.R., Drummond L.O., Teixeira Jr. M., Rodrigues M.T.,

Carnaval A.C. (2017). Biogeographic links between southern Atlantic Forest and western South

America: rediscovery, re-description, and phylogenetic relationships of two rare montane anole lizards from Brazil. Molecular Phylogenetics and Evolution, 113: 49-58.

Chapter 3: Prates I., Rivera D., Rodrigues M.T., Carnaval, A.C. (2016). A mid‐Pleistocene rainforest corridor enabled synchronous invasions of the Atlantic Forest by Amazonian anole lizards. Molecular Ecology, 25: 5174–5186.

x

Table of Contents Page

Introduction 1

Chapter 1: Phylogenetic relationships of Amazonian anole lizards (Anolis: Dactyloa):

taxonomic implications, new insights about phenotypic evolution and the timing of

diversification 7

Chapter 2: Biogeographic links between southern Atlantic Forest and western South

America: rediscovery, redescription, and phylogenetic relationships of two rare

montane anole lizards from Brazil 25

Chapter 3: A mid-Pleistocene rainforest corridor enabled synchronous invasions of the

Atlantic Forest by Amazonian anole lizards 46

Chapter 4: Genomic divergence and local adaptation during range expansions in

widespread South American anoles 67

Tables 83

Figures 91

Appendices 109

Literature Cited 129

xi

List of Tables Page

Table 1.1: Median values and 95% highest posterior density intervals (in millions of years) of log-normally distributed calibration priors applied in dating analyses, based on fossil data. Settings for calibration prior mean, standard deviation and offset are provided.

MRCA = most recent common ancestor. 83

Table 3.1: Posterior estimates of the time of divergence (T, in millions of years) between lizards in the Atlantic Forest (AF) and eastern Amazonia (EAm), the time of population bottlenecks in the Atlantic Forest, and effective population sizes (Ne, in millions of individuals), based on best-fit scenarios in DIYABC analyses. 95% credibility intervals

(CI) are indicated. 84

Table 4.1: Posterior estimates from G-PhoCS analyses of Anolis ortonii and A. punctatus.

Model parameters are the time of coalescence between populations (T, in millions of years), effective population sizes (N, in millions of individuals), and migration rates (M, the proportion of individuals in one population that arrived from another population in each generation) between Atlantic Forest lizards (af) and their most closely-related

Amazonian population (am). Estimates for ancestral populations (anc) and western

Amazonian populations of A. punctatus (wamz) are also indicated. CI denotes 95% credibility intervals. 85

Table 4.2: Candidate loci recovered by genome-environment association analyses using

LFMM for Anolis ortonii and A. punctatus. Only genes that successfully blasted against protein-coding regions of the genome of A. carolinensis and were successfully annotated

xii for gene ontologies are shown. Gene ontology terms (GO names) for biological activities at the molecular (F) and cell or organism level (P) are indicated. 86

xiii

List of Figures Page

Figure 1.1: Dewlap coloration patterns of our focal Amazonian anole lizards, Anolis punctatus (A), A. philopunctatus (B), A. phyllorhinus (C), and A. dissimilis (D). 91

Figure 1.2: Collection sites of new Dactyloa clade samples included in this study.

Orange: Anolis punctatus with orange dewlaps. White: A. punctatus with white dewlaps.

Purple: A. philopunctatus. Red: A. phyllorhinus. Green: A. dissimilis. Blue: A. transversalis. Abbreviations of sampled states in Brazil are indicated. 1. Fazenda

Experimental Catuaba, Acre (AC): A. dissimilis UFAC0075, UFAC0084, UFAC0089, A. punctatus MTR28593, A. transversalis MTR28583; 2. Serra do Divisor National Park,

AC: A. punctatus MTR28048; 3. Porto Velho, Rondônia (RO), A. punctatus H1907; 4.

Pacaás Novos National Park, RO: A. punctatus MTR26005; 5. Itapuru, Amazonas (AM):

A. punctatus MTR18550; 6. Coari, AM: A. punctatus MPEG26911; 7. Novo Aripuanã,

AM: A. phyllorhinus MTR14041; 8. Colniza, (MT): A. phyllorhinus

MTR977398, MTR977423, MTR977628, MTR977664, A. punctatus MTR967985; 9.

Jacareacanga, Pará (PA): A. phyllorhinus PJD002; 10. Manaus, AM: A. philopunctatus

MTR21474; 11. Presidente Figueiredo, AM: A. philopunctatus Galo132; 12. Urucará,

AM: A. punctatus MPEG29316; 13. Oriximiná, PA: A. punctatus MPEG24758; 14.

Juruti, PA: A. punctatus MPEG28489; 15. Afuá, PA: A. punctatus MPEG29591; 16.

Camacan, Bahia (BA): A. punctatus MTR16109; 17. Linhares, Espírito Santo (ES): A. punctatus MTR12509. 92

Figure 1.3: Dewlap coloration of preserved Anolis punctatus (top) and A. philopunctatus.

xiv

Both patterns remain promptly discernible even after at 27 years of fixation. 94

Figure 1.4: Coalescent-based phylogeny of the Dactyloa clade of Anolis inferred using

*BEAST. Posterior probabilities = 1 are indicated with an asterisk. Species groups follow

Castañeda and de Queiroz (2013). Picture credits: Julián Velasco (A. calimae). 95

Figure 1.5: Phylogeny of the Dactyloa clade of Anolis based on a concatenated dataset.

Bayesian (MrBayes) and maximum likelihood (Garli) analyses recovered the same topology. Node values represent posterior probability/bootstrap support values. Posterior probabilities = 1 and bootstrap values = 100 are indicated with an asterisk. Species groups follow Castañeda and de Queiroz (2013). Picture credits: Alejandro Arteaga, Tropical

Herping (A. proboscis), Pedro Peloso (A. transversalis), Renato Recoder (A. punctatus),

Bret Whitney (A. phyllorhinus). 96

Figure 1.6: Divergence times between pleurodont iguanian lizards, based on five nuclear markers (4,082 base pairs) and inferred using BEAST. Red circles denote calibrated nodes. Posterior probabilities = 1 are indicated with an asterisk. Bars represent the 95% highest posterior densities (HPD). Picture credits: Pedro Peloso (Anolis trachyderma). 98

Figure 2.1: Coloration in life of Anolis nasofrontalis (A, B) and A. pseudotigrinus (C, D).

In A, inset shows the black throat lining of A. nasofrontalis. Photographed specimens are females. 99

Figure 2.2: Phylogenetic relationships and divergence times between species in the

Dactyloa clade of Anolis inferred using BEAST. Asterisks denote posterior probabilities > 0.95. 100

xv

Figure 2.3: Head in dorsal and ventral views of Anolis nasofrontalis (A, B) and A. pseudotigrinus (C, D). 101

Figure 2.4: Geographic distribution of confirmed and purported members of the neblininus species series within the Dactyloa clade of Anolis. Anolis bellipeniculus: Cerro

Yaví, Amazonas, Venezuela (Myers and Donelly, 1996); A. calimae: San Antonio,

Television Tower Mountain, Valle del Cauca, Colombia, and Lake Calima Dam, Daríen,

Valle del Cauca, Colombia (Ayala et al., 1983); A. carlostoddi: Abacapa-tepui, Chimantá massif, Bolivar, Venezuela (Williams et al., 1996); A. dissimilis: Fazenda Experimental

Catuaba, Senador Guiomard, Acre, Brazil (Melo-Sampaio et al., 2013; Prates et al.,

2015), Parque Estadual Chandless, Manoel Urbano, Acre, Brazil (Freitas et al., 2013),

San Martín, Lower Urubamba Region, (Icochea et al., 2001), and Itahuania, upper

Rio Madre de Dios, Madre de Dios, Peru (Williams, 1965); Anolis nasofrontalis and A. pseudotigrinus: Santa Teresa, Espírito Santo, Brazil (this study; Myers and Carvalho,

1945; Williams and Vanzolini, 1980); A. neblininus: Cerro de la Neblina, Amazonas,

Venezuela (Myers et al., 1993); A. peruensis: Esperanza, Amazonas, Peru (Poe et al.,

2015); A. williamsmittermeierorum: Venceremos, San Martin, Peru (Poe and Yañez-

Miranda, 2007). For the last two taxa, a single point is depicted on map due to close proximity of collection sites. The inset presents a schematic map of South America around 10-12 mya (based on Lundberg et al. 1998), illustrating the approximate locality of the Chapare buttress, a land bridge that connected the central Andes to the western edge of the Brazilian Shield during the Miocene. 102

Figure 3.1: Demographic scenarios tested with coalescent simulations and ABC. A)

xvi

Northeastern vicariance: an ancestral population occurring in both eastern Amazonia

(EAm) and in the Atlantic Forest (AF) is split due to expansion of open and dry domains in presently dry northeastern South America. B) Northeastern dispersal: a founder population colonizes the AF from EAm, with population expansion following a bottleneck. Scenarios C (southwestern vicariance) and D (southwestern dispersal) are similar to A and B respectively, with the difference that the southwestern (instead of eastern) Amazonia acts as the source of dispersal into the Atlantic Forest. T = time. E)

Map showing the three regions considered in this study. 104

Figure 3.2: Sampled sites and coalescent-based phylogeny of Anolis punctatus (A), A. ortonii (B), and Polychrus marmoratus (C) based on a multi-locus dataset, depicting the relationships between putative independently evolving lineages as inferred based on bGMYC. Colors correspond to the spatial groups used in analyses of alternative demographic scenarios in DIYABC, as follows: blue, southwestern Amazonia; orange, eastern Amazonia; red; Atlantic Forest. Brazilian state acronyms on maps are as follows:

AC, Acre; AL, Alagoas; AM, Amazonas; BA, Bahia; ES, Espírito Santo; MA, Maranhão;

MG, Minas Gerais; MT, Mato Grosso; PA, Pará; RJ, Rio de Janeiro; RO, Rondônia; SE,

Sergipe; SP, São Paulo; TO, Tocantins. 105

Figure 4.1: Historical demographic models and parameters estimated using G-PhoCS for

Anolis ortonii (top) and A. punctatus. N, effective population sizes; T, time of population coalescence; af, Atlantic Forest population; am, Amazonian population (in A. ortonii); anc, ancestral populations; eam, eastern Amazonian population (in A. punctatus); weam, western Amazonian populations (in A. punctatus). 106

xvii

Figure 4.2: Patterns of genetic structure inferred using sNMF for Anolis ortonii (top) and

A. punctatus. A population tree of A. punctatus estimated with SNAPP is also indicated. 107

Figure 4.3: Ontology annotations of candidate loci recovered by genome-environment association analyses performed using LFMM for Anolis ortonii (pies on the left) and A. punctatus. Activities at the level of molecular function (top) and biological process at the level of cell or organism are indicated. For simplicity, pie slices that correspond to minor categories that are not shared by the two anole species are indicated in gray. 108

xviii

Introduction

Shifts in the geographic distribution of habitats over time can promote dispersal and vicariance, thereby influencing large-scale biogeographic patterns and ecological processes

(Graham et al., 2004; Santos et al., 2009). The disruption of environmental barriers, for instance, provides opportunities for species colonization of new geographic regions (Petren et al., 2005).

Such range expansions can lead organisms to encounter and interact with new species and environments, which may trigger evolutionary radiations (Hughes and Eastwood, 2006). The establishment of corridors of suitable habitat across previously disjunct yet ecologically similar regions has been widely linked to climate change over time (Brown et al., 2013; Kozak and

Wiens, 2006). Such climate-mediated habitat shifts may have played a key role in the assembly of tropical biotas, including those of Amazonia and the Atlantic Forest in South America, two of the most diverse ecosystems on Earth. The history of biogeographic associations between tropical rainforest habitats in response to former climatic shifts, as well as their consequences for the distribution and genetic composition of forest organisms, is the theme of this doctoral dissertation.

Amazonian and Atlantic rainforests are currently separated by the savannas and xeric scrublands of the Caatinga, Cerrado, and Chaco domains, yet there is substantial evidence that these forests, and their communities, have been linked through time. Historical biotic interchange is supported by present-day floristic similarity (Bigarella et al., 1975; Santos et al., 2007) and disjunct distribution patterns in an expanse of and plant species (e.g., Fiaschi and Pirani,

2009; Gehara et al., 2014; Ribeiro-Júnior, 2015; Rocha et al., 2015). Sister relationships between clades restricted to either Amazonia or the Atlantic Forest suggest that in situ diversification following dispersal has contributed to the high levels of endemism seen in both systems (e.g.,

1

Batalha-Filho et al., 2013; Costa, 2003; Fouquet et al., 2012a, b). Importantly, despite the potential significance of biotic interchange during the assembly of ecological communities in

Amazonian and the Atlantic Forest, our knowledge about the timing, frequency, and magnitude of such presumed biogeographic connections is still limited – a gap that this dissertation aims to help fill.

Environmental variation across the geographic space is known to constrain species range limits through strong effects on the physiological and behavioral performance of organisms, ultimately influencing biogeographic patterns and ecological processes (Ghalambor et al., 2006;

Navas, 2002). As a result, it has been suggested that local adaptation plays a major role in the establishment of organisms in new environments (Hohenlohe et al., 2010; Huey et al., 2000).

However, experimental studies support that physiological tolerances are generally highly conserved within species over wide environmental gradients (e.g., Crowley, 1985; Hertz et al.,

1983; John-Alder et al., 1989; Van Damme et al., 1989), while the potential for local adaptation is greatly limited by the homogenizing effects of gene flow between populations (Lenormand,

2002; Bridle and Vines, 2007; Sexton et al., 2014). To assess whether and how local adaptation plays a role during the establishment of species in new environments, this research aims to help uncover how ecological gradients affect the genetic makeup of organisms, as well as how historical demographic factors - such as gene flow - may constraint population differentiation.

Because local adaptation has been implicated in genetic divergence and diversification (Nosil,

2012), this work can improve our understanding of the process of local adaptation and associated biological diversification across ecological gradients (Rellstab et al., 2015).

To perform biogeographic, historical demographic, and comparative genomic inference, the studies presented in this dissertation employ genetic information from South American

2

species of true anoles (Anolis, Dactyloidae) and, to a lesser extent, bush anoles (Polychrus,

Polychrotidae). These lizards are well suited for tackling the main goals of this investigation.

Because several anole species are strongly associated with closed-canopy wet forests, their phylogenetic and demographic histories can provide insights about former shifts in forest distribution (e.g., Carnaval et al., 2009). Species such as Anolis fuscoauratus, Anolis ortonii,

Anolis punctatus, and Polychrus marmoratus occur disjunctively in Amazonia and the Atlantic

Forest, but not in the open and dry ecosystems that separate these two forest blocks (Ribeiro-

Júnior, 2015). Other species, such as the southern Atlantic Forest montane endemics Anolis nasofrontalis and Anolis pseudotigrinus, have unclear phylogenetic affinities and peculiar morphologies, being more similar to anoles from other distant South American mountains, such as the Andes and Guianan highlands (Williams, 1976, 1992). Because the distribution and phenotypic attributes of these anoles may be indicative of historical biogeographic links among regions, they represent a promising system to investigate former shifts in the distribution of

South American rainforests.

The ecology and thermal biology of anoles also make them a suitable group for studies of how environmental variation across space affects genomic variation. Several South American forest anoles do not rely on behavioral thermoregulation (Vitt et al., 2002, 2003), which points to a role of genetically determined physiological traits in shaping ecological tolerances. Studies on

Caribbean Anolis suggest that abiotic factors strongly affect behavioral and physiological performance and constrain species ranges in this group (Huey et al., 1983; Knouft et al., 2006;

Losos, 2009), while thermal niches may evolve rapidly in response to environmental regimes

(Kolbe et al., 2012; Leal and Gunderson, 2012; Muñoz et al., 2014). The recent sequencing and annotation of the genome of the green anole, Anolis carolinensis (Alföldi et al., 2011), has the

3

potential to support comparative genomic studies in this group. Importantly, while

Caribbean anoles are central to studies of adaptive radiation and evolutionary ecology, we know very little about the biology and evolution of mainland species, which effectively correspond to most of the taxonomic and ecological diversity of the anole family (Losos, 2009).

In the Neotropical region, lack of basic knowledge about species boundaries, evolutionary relationships, and the timing of diversification strongly constrains inferences about historical biogeography and the drivers of diversification (Angulo and Icochea, 2010; Smith et al., 2013). As a result, my dissertation begins with a revisionary phylogenetic study of poorly- known rainforest anole taxa. Chapter 1 provides a multi-locus phylogeny of a highly diverse clade within Anolis, known as the Dactyloa clade, to clarify the taxonomic status, geographic distribution, and morphological variation of understudied Amazonian and Atlantic Forest species that I have sampled through my own fieldwork and collaborations. Based on a critical review of the fossil record of the pleurodont iguanian lizard clade, that study also provides a temporal framework for the diversification of South American lizards in the genera Anolis and Polychrus, therefore yielding much-needed estimates of divergence times and mutation rates for my subsequent phylogeographic and population genetic analyses. A manuscript describing the results of this work was published in Molecular Phylogenetics and Evolution (Prates et al.,

2015).

In Chapter 2, I further my evolutionary studies of Anolis to examine historical biogeographical relationships between South American regions. In this study, I incorporate newly obtained samples of narrowly-distributed taxa, including the rare A. nasofrontalis and A. pseudotigrinus, which have been undocumented for more than 40 years. Based on molecular data, I find support for Miocene biotic interchange between the southern Atlantic Forest, western

4

Amazonia, Andean Yungas, and Guianan highlands. My morphological examinations of A. nasofrontalis, A. pseudotigrinus, and their closest relatives point to a history of recurrent evolution of a twig anole-like phenotype in mainland anoles, in convergence with Caribbean taxa and in apparent association with montane settings. To support future biodiversity inventories in the Atlantic Forest, I also provide a taxonomic redescription of A. nasofrontalis and A. pseudotigrinus. The results from this study were published in Molecular Phylogenetics and

Evolution (Prates et al., 2017).

Chapters 1 and 2 provided the phylogenetic, taxonomic, and temporal context for my third dissertation chapter. In Chapter 3, I employ sequence data at multiple loci in a comparative phylogeographic study of three co-distributed arboreal lizards (Anolis punctatus, A. ortonii, and

Polychrus marmoratus) to investigate the timing, magnitude, and spatial routes of proposed former corridors between Amazonia and the Atlantic Forest. By testing alternative historical demographic scenarios of colonization and vicariance using coalescent simulations and

Approximate Bayesian Computation (ABC), I find evidence that climatic fluctuations during the

Pleistocene provided key opportunities for synchronous colonizations by co-distributed forest taxa through the cessation of environmental barriers in eastern South America. Interestingly, evidence of species-specific responses strengthens assertions that biological attributes play a role in responses to shared environmental change. This work has been published in Molecular

Ecology (Prates et al., 2016b).

Uncovering this history of range expansions in widely-distributed South American anoles provided the foundation for my fourth dissertation chapter. In Chapter 4, I perform a comparative population genomic study of Anolis ortonii and A. punctatus, which according to my Chapter 3 colonized the Atlantic Rainforest from eastern Amazonia. Based on thousands of restriction

5

enzyme-associated DNA markers, I assess spatial congruency between genetic and environmental breaks and revisit the history of range expansions under a coalescent framework, evaluating levels of gene flow among populations – a force that could oppose to local adaptation.

Moreover, to assess whether species occurrence in diverse environments have relied on local adaptation, I implement genome-environment association analyses, testing for correlations between allele frequencies with gradients of temperature and precipitation across each species’ range. The results point to adaptive genomic differentiation during range expansions across heterogeneous landscapes in widely distributed species. This genetic divergence may have been favored by large effective population sizes and low migration rates across regions, pointing to interactions between natural selection and the demographic trajectory of populations in the process of adaptation. A manuscript reporting the findings of my Chapter 4 is now being prepared for submission to Evolution.

6

CHAPTER 1: Phylogenetic relationships of Amazonian anole lizards (Anolis: Dactyloa):

taxonomic implications, new insights about phenotypic evolution and the timing of

diversification1

Introduction

Caribbean anole lizards have been central to studies of adaptive radiation and evolutionary ecology for more than half a century (Losos, 2009). However, most of the diversity within the anole genus (Anolis, family Dactyloidae) is, in fact, continental (Nicholson et al.,

2012). In contrast to the island forms, little is known about the biology and evolution of mainland anoles. In the species-rich clade within Anolis known as Dactyloa (83 recognized species; Castañeda et. al., 2014), unclear phylogenetic affinities limit inferences about species boundaries, morphological evolution, and biogeographic patterns. We help fill this gap by combining morphological and DNA sequence data from four species of previously unsampled or undersampled Amazonian anoles. Based on new information of Anolis punctatus, Anolis philopunctatus, Anolis phyllorhinus, and Anolis dissimilis, we also provide an updated molecular phylogeny of Dactyloa anoles.

The Dactyloa anole radiation is composed of five main clades that define groups of species often referred to as ‘species series’ (Castañeda and de Queiroz, 2013). Most Amazonian

Dactyloa species are currently assigned to the punctatus series, which includes ca. 20 taxa, some of which exhibit wide ranges in South America (Castañeda and de Queiroz, 2013). Anolis punctatus Daudin 1802, the spotted or green Amazon anole (Fig. 1.1a), is one such example that occurs throughout the Amazon basin, the Andean foothills, the Guiana Shield, and the Atlantic

1 Published as: Prates I., Rodrigues M.T., Melo-Sampaio P.R., Carnaval A.C. (2015). Phylogenetic relationships of Amazonian anole lizards (Dactyloa): taxonomic implications, new insights about phenotypic evolution and the timing of diversification. Molecular Phylogenetics and Evolution, 82: 258-268. 7

Forest of coastal Brazil. A previous assessment of population genetic structure in this morphologically conserved lizard was limited to four localities (Glor et al., 2001). To assess the biological and taxonomic implications of genetic variation within A. punctatus, we combine observations of color variation of a key sexual signal, the anole dewlap, with new sequence data from a representative portion of the species’ range.

Molecular evidence may improve the delimitation of species within the punctatus group, which has often been challenging (Ugueto et al., 2007; Williams, 1982). In this study, we focus particularly on A. philopunctatus Rodrigues 1988 (Fig. 1.1b), a name given to anole populations restricted to central Amazonia that are morphologically distinguishable from A. punctatus solely based on the coloration of the dewlap in male specimens. The dewlap of A. punctatus presents rows of small whitish scales on an orange background, while that of A. philopunctatus has large scattered black spots on an orange background (Fig. 1.1). The typical orange dewlap of A. punctatus has been described as nearly invariable throughout this species’ distribution (Avila-

Pires, 1995; Rodrigues, 1988; Williams, 1982). However, at least one population in southern

Brazilian Amazonia presents white dewlaps (Rodrigues et al., 2002). Because there are no records of A. punctatus and A. philopunctatus occurring in sympatry and females of the two species cannot be distinguished based on their extremely reduced dewlaps, the validity of A. philopunctatus has been questioned (Avila-Pires, 1995). Making matters more complex, a recent phylogenetic assessment based on morphological data of A. philopunctatus recovered this species not within the punctatus species series but closely related to the latifrons series, albeit with low support (Castañeda and de Queiroz, 2013). Using sequence data, we assess the phylogenetic placement of A. philopunctatus relative to A. punctatus, asking whether dewlap color pattern is a robust diagnostic character in the face of genetic variation within A. punctatus.

8

Limited phylogenetic information also constrains inferences regarding morphological evolution within the anole radiation, as is the case of the poorly understood rostral proboscis observed in A. phyllorhinus Myers and Carvalho 1945 (Fig. 1.1c), Anolis laevis Cope 1876, and

Anolis proboscis Peters and Orces 1956. Based on the possession of the proboscis, Williams

(1979) grouped these three taxa in the laevis species group. However, they present highly disjunct distributions, as A. laevis occupies the eastern Andean foothills in Peru, A. proboscis occurs in the western slopes of the Ecuadorian Andes, and A. phyllorhinus is found in southern

Amazonian lowlands in Brazil. To date, A. proboscis has been the only proboscis-bearing anole studied under a phylogenetic framework. Based on morphological characters, it has been consistently grouped with taxa in the heterodermus species series (Castañeda and de Queiroz,

2013; Nicholson et al., 2012; Poe, 2004; Poe et al., 2012). By contrast, morphological examinations suggested A. phyllorhinus as related to the punctatus series (Rodrigues et al., 2002;

Yánez-Muñoz et al., 2010). If A. phyllorhinus is in fact more closely related to A. punctatus than to the remaining proboscis-bearing species, anole rostral appendages have evolved at least twice.

By uncovering the phylogenetic affinities of A. phyllorhinus, we evaluate the hypothesis of independent nasal appendage evolution within the anole clade.

The paucity of anole studies in South America also limits our knowledge about species distributions. For almost 50 years, Anolis dissimilis Williams 1965 (Fig. 1.1d) was known solely from its type locality in the upper Madre de Dios River (Peru). Williams (1965, 1974, 1982) has suggested that this ‘odd anole’ was related to the punctatus species series. This assignment was supported by phylogenetic analyses incorporating morphological evidence, but the low support of the resulting trees and conflicting topologies rendered it tentative (Castañeda and de Queiroz,

2013). Recently, A. dissimilis was found in Brazil (Freitas et al., 2013). We found this species to

9

be common in a southwestern Amazonian site (Melo-Sampaio et al., 2013), where it is associated with clumps of open bamboo forest (Guadua spp.) within dense rainforest. By incorporating A. dissimilis in our analysis, we test for its presumed affinity with the punctatus species series.

Finally, to lay the groundwork for forthcoming phylogeographic studies, we combine multi-locus sequence data with a critical review of the available fossil information, providing a new dated phylogeny of the pleurodont iguanian clade that includes representatives of both the

Dactyloa and Norops clades within Anolis, as well as bush anoles (Polychrus). By performing improved estimations of divergence times, we provide a temporal framework for Amazonian anole diversification.

Our analysis incorporates molecular data generated by a recent phylogenetic assessment of Dactyloa (Castañeda and de Queiroz, 2011), yet we do not reexamine the group’s systematics beyond our target Amazonian taxa. Instead, we refer to the much more extensive work of

Castañeda and de Queiroz (2013).

Material and Methods

Sampling of molecular data

We generated DNA sequences of six Anolis phyllorhinus specimens (from three collection sites, Fig. 1.2), three A. dissimilis (one site), two A. philopunctatus (two sites), 12 A. punctatus with orange dewlaps (10 Amazonian and two Atlantic Forest sites), one Amazonian A. punctatus with a white dewlap, one specimen of Anolis transversalis, one of each of four

Amazonian species of Norops clade anoles (Anolis fuscoauratus, Anolis ortonii, Anolis tandai and Anolis trachyderma), and one individual of the bush anole Polychrus liogaster

10

(Polychrotidae). All samples were collected in Brazil (see Appendix 1.1 for a list of voucher numbers, locality information and Genbank accession numbers). Because the examination of preserved specimens (including the type series of A. philopunctatus) demonstrated that dewlap patterns remain promptly discernible even after 27 years of fixation (Fig. 1.3), we only sampled tissues from localities where the dewlap pattern is known, based on direct observation of live or preserved specimens.

For phylogenetic inference within the Dactyloa clade of Anolis, we matched available datasets (Castañeda and de Queiroz, 2011) with new sequences of the mitochondrial gene NADH dehydrogenase subunit 2 (ND2) and the flanking tryptophan transfer RNA (tRNA-Trp) gene, following Jezkova et al. (2009). Sequences obtained in Genbank often included four additional tRNAs flanking the ND2 gene (tRNA-Ala, tRNA-Asx, tRNA-Cys, tRNA-Tyr) (Castañeda and de Queiroz, 2011), which were also used in our final alignments (see below). Additionally, we generated sequences of the nuclear recombination-activating gene 1 (RAG1), as per Gartner et al. (2013). For divergence time estimation within pleurodont iguanians, we sampled the nuclear genes BTB and CNC homology 1 (BACH1), dynein axonemal heavy chain 3 (DNAH3), megakaryoblastic leukemia 1 (MKL1), nerve growth factor beta polypeptide (NGFB) and synuclein alpha interacting protein (SNCAIP), as per Townsend et al. (2008, 2011). Sequences were edited and aligned using Geneious Pro 6 (Biomatters, Auckland).

We used Geneious’ plugin Find Heterozygotes with a 0.90 overlap threshold to identify heterozygous positions. For the coalescent-based phylogenetic analyses (see below), we ran nuclear sequences through PHASE 2.1.1 (Stephens and Donnelly, 2003) to estimate the haplotypic phase of heterozygotes, after preparation of input files in SeqPHASE (Flot, 2010). We ran PHASE ten independent times, using a 0.90 probability threshold and a parent-independent

11

mutation model. Models of nucleotide evolution and best-fit partition schemes, including partitions by codon position, were determined with Partition Finder 1.1.1 (Lanfear et al., 2012), implementing PhyML for likelihood estimation (Guindon and Gascuel, 2003) and the Akaike information criterion for model selection (Akaike, 1974). The short length of the five tRNAs markers (~70 bp) prevented Partition Finder from properly estimating substitution parameters for those regions. As a result, we treated all tRNAs (397 base pairs total) as a single partition. A concatenated dataset was generated in Sequence Matrix (Vaidya et al., 2011).

Inferring phylogenetic relationships

For phylogenetic analyses, we combined our sequences with a subset of the Dactyloa clade molecular dataset generated by Castañeda and de Queiroz (2011). We only included sequences from specimens with traceable voucher numbers. To meet the requirements of the coalescent-based analyses, we only used data for species that had both the mitochondrial and the nuclear fragments sequenced. Also, we did not combine gene fragments of different individuals in chimeric sequences. These criteria ensured that the exact same dataset was used for both coalescent-based and concatenated analyses (see below). Our final dataset was composed of 77 specimens of 33 Dactyloa anole species.

Phylogenetic inference based on multiple loci traditionally involves a concatenation strategy, in which different markers are juxtaposed in a single alignment with the implicit assumption that the different genes have a common genealogy. Yet, gene trees are often discordant and fail to reflect the real relationships among species because stochasticity in the coalescent process may prevent complete fixation and reciprocal monophyly of alleles (Kubatko and Degnan, 2007). By incorporating population parameters such as ancestral population sizes

12

and divergence times, coalescent-based species tree methods account for ancestral polymorphism and incomplete lineage sorting (Edwards, 2009). We implemented a coalescent-based approach using the *BEAST tool (Heled and Drummond, 2010) of the BEAST 1.8 package (Drummond et al., 2012), applying five independent runs of 100 million generations each and sampling every

10,000 steps. We manually edited the xml files to ensure a single mitochondrial tree while implementing distinct substitution models for ND2 and the tRNA genes, and used phased data for the nuclear RAG1 gene. Because *BEAST analyses consistently failed to reach stationarity when partitioning protein-coding genes by codon position, suggesting model over- parameterization, we implemented gene partitions only in our coalescent-based analyses.

*BEAST requires a priori assignment of individuals to species; because concatenated analyses and individual gene trees consistently found Anolis philopunctatus to be nested within A. punctatus (see Results), we treated both as a single species (A. punctatus) to avoid specifying paraphyletic taxa for coalescent-based phylogenetic inference.

We compared *BEAST’s results to trees generated under a concatenated approach, using both maximum likelihood and Bayesian inference. We ran maximum likelihood analyses in Garli

2.0 (Zwickl, 2006), with 100 replicates for tree search and 1,000 bootstraps to assess clade support. We summarized bootstraps on the best resulting tree with the SumTrees tool of the

DendroPy 3.12 Python package (Sukumaran and Holder, 2010). Additionally, we estimated a

Bayesian tree with MrBayes 3.2.1 (Ronquist et al., 2012) through three independent runs and four Markov chains of 20 million generations each, sampling every 1,000 steps. In both MrBayes and Garli analyses, we partitioned protein-coding genes by codon position as indicated by

Partition Finder.

For the Bayesian analyses (*BEAST and MrBayes), we assessed convergence and

13

stationarity of model parameters using Tracer 1.5 (Drummond et al., 2012), combined runs in

LogCombiner 1.8 (with a 25% burn-in), and summarized a maximum clade credibility tree in

TreeAnnotator 1.8 (Drummond et al., 2012). In all analyses, we unlinked parameters of substitution rates and nucleotide frequencies between partitions. The resulting topologies were visualized in FigTree 1.4 (available from http://tree.bio.ed.ac.uk/software/figtree/).

For descriptive purposes, we also estimated Tamura-Nei corrected pairwise genetic distances (Tamura and Nei, 1993) for the mitochondrial DNA fragment of the newly sampled

Dactyloa specimens, using the APE 3.1 package (Paradis et al., 2004) of the R 3.0.2 platform (R

Core Team, 2017).

Divergence time estimation

A recent phylogenetic assessment of Anolis estimated divergence times between

Amazonian species and their closest relatives (Nicholson et al., 2012), yet with a few caveats.

For instance, the phylogenetic placement of the semi-fossils used for node calibration has been uncertain (Castañeda et al., 2014; de Queiroz et al., 1998; Lazell, 1965). Moreover, the divergence time estimates presented by Nicholson et al. (2012) were based on a single mitochondrial marker, yet mitochondrial genes are known to mislead deep divergence time estimation due to substitution saturation (Brandley et al., 2011; Lukoschek et al., 2012; Mulcahy et al., 2012), sometimes resulting in 3 to 10-fold overestimations (Zheng et al., 2011). To improve estimates of divergence times for Amazonian anoles, we built a dated phylogeny for the pleurodont iguanian clade (sensu Wiens et al., 2012) using five nuclear genes and well-known fossils for calibration points. For that, we combined newly generated sequences of Amazonian species of Anolis within both the Dactyloa and Norops clades, as well of Polychrus

14

(Polychrotidae), with published sequences of 12 other pleurodont lizard taxa, emphasizing South

American genera (see Appendix 1.2 for taxa, voucher and Genbank accession numbers).

We used three fossils for node calibration: 1) Saichangurvel davidsoni, from the late

Cretaceous (Conrad and Norell, 2007), calibrated the node defining Pleurodonta, 2) Suzanniwana sp. (Corytophanidae), from the early Eocene (Smith, 2009, 2011), was placed at the stem of the clade defined by Corytophanes and Basiliscus, and 3) Afairiguana avius (Leiosauridae), from the early Eocene (Conrad et al., 2007), was placed at the stem of the clade defined by Leiosaurus and Urostrophus. In the latter case, we conservatively treated A. avius as a stem leiosaurid because Conrad et al. (2007) found the relationships between A. avius and extant leiosaurids to be unresolved. Following best practices for divergence time estimation (Parham et al., 2012), we did not incorporate most of the pleurodont fossils used as calibration points to date (e.g.

Towsend et al., 2011) due to the lack of phylogenetic analyses or explicit synapomorphies for fossil placement (e.g., Holman, 1972, 1987; Smith, 2006; Yatkola, 1976). In fact, some of those fossils cannot be unambiguously assigned to any iguanian group (Sullivan and Holman, 1996).

Because setting maximum age bounds for calibration priors is challenging in the face of the incompleteness and uncertainty of the fossil record (Ho and Phillips, 2009), we conservatively assigned long tails to the probability distribution of lognormally distributed calibration priors.

Calibration prior settings, including median values and the 95% highest posterior density intervals, are presented in Table 1.1.

We performed simultaneous phylogenetic reconstruction and divergence time estimation by implementing an uncorrelated lognormal relaxed clock (Drummond et al., 2006) under a concatenation approach in BEAST. We implemented a uniform prior distribution (interval = 0-1) to the mean rate of the molecular clock (ucld.mean parameter), while using default settings for

15

the parameters relative to substitution rates, nucleotide frequencies, and the Yule tree prior.

Because the use of codon partitions consistently prevented proper Markov chain mixing, we only partitioned the dataset by gene when running the dating analyses. We ran five independent chains of 100 million steps, sampling every 10,000 steps. After assessing stationarity and convergence of model parameters in Tracer (Drummond et al., 2012), we applied a 25% burn-in, combined the runs, and summarized results into a maximum clade credibility tree as described above.

Results

Phylogenetic relationships of Amazonian Dactyloa

Both the coalescent-based (Fig. 1.4) and the concatenated (Fig. 1.5) analyses recovered six main clades within Dactyloa. Five of them correspond to the previously recognized aequatorialis, heterodermus, latifrons, punctatus and roquet species series, with nearly the same species composition as listed by Castañeda and de Queiroz (2013) (exceptions were the phylogenetic placement of Anolis fitchi and A. philopunctatus, see below). The sixth recovered clade is composed of Anolis calimae, Anolis neblininus, and the newly sampled A. dissimilis.

This clade was highly supported in both the coalescent-based (*BEAST) and Bayesian concatenated (MrBayes) analyses (posterior probabilities ≥ .98), yet weakly supported in the maximum likelihood (Garli) tree (bootstrap support = 68%). The two concatenated analyses

(MrBayes, Garli) recovered the same topology for the entire Dactyloa clade phylogeny.

Relationships among these six major clades were poorly supported and inconsistent across analyses. For instance, the punctatus species series was inferred as the sister of the remaining Dactyloa in the coalescent-based phylogeny (Fig. 1.4), but not in the trees generated

16

through a concatenation approach (Fig. 1.5). Anolis fitchi, the only species assigned to different main clades across our analyses, was recovered with low support as the sister of the clade composed by A. dissimilis, A. neblininus, and A. calimae in the coalescent-based phylogeny (Fig.

1.4). Yet, in the concatenated analyses, A. fitchi was placed as closely related to the aequatorialis group (Fig. 1.5), a relationship previously recovered by Castañeda and de Queiroz (2013).

New sequence data provide important information about relationships within the Anolis punctatus group. All samples of A. phyllorhinus were deeply nested within this clade, with high support (Fig. 1.5). This species was recovered as A. punctatus’ sister taxon in all analyses, with pairwise corrected genetic distances between A. phyllorhinus and A. punctatus ranging 19-25%.

A clade comprised of samples of Anolis punctatus, which have orange dewlaps, and of A. philopunctatus, which have spotted dewlaps, was consistently highly supported (Fig. 1.5).

Although our two samples of A. philopunctatus composed a clade, the analyses found this species to be nested within A. punctatus. Corrected pairwise genetic distances between A. philopunctatus samples and closely related A. punctatus were relatively small, between 3.4 and

5.1%. By contrast, Amazonian populations of A. punctatus presenting the orange dewlap pattern often exhibited higher pairwise genetic distances, from 3.4 % to 12%. Relationships within the A. punctatus + A. philopunctatus clade were poorly supported. Interestingly, those samples of A. punctatus collected in the Atlantic Forest (Camacan and Linhares localities) were nested among

Amazonian ones (Fig. 1.5).

Similar to the Anolis philopunctatus individuals, the only sample of A. punctatus with a white dewlap (MTR967985) was recovered nested among A. punctatus specimens exhibiting orange dewlaps (Fig. 1.5). The lowest genetic distance separating this individual from a sample presenting the typical A. punctatus pattern was 6.4%.

17

Divergence times between Amazonian lizards

Dating analyses suggest that lineage diversification within the pleurodont iguanian clade dates back to the late Cretaceous (Fig. 1.6). We estimated the most recent common ancestor

(MRCA) of sampled pleurodonts to have lived at approximately 82 million years ago (Mya)

(median value; 95% of the highest posterior density [HPD] = 71-99 Mya). On the other hand, some of the divergences between Neotropical genera are as recent as the Middle Miocene (e.g.,

Basiliscus and Corytophanes, Urostrophus and Leiosaurus, Morunasaurus and Enyalioides). Our analyses recovered the anole family (Dactyloidae) as the sister of all other sampled pleurodonts.

Among extant anoles, our analyses suggest that Dactyloa diverged from the MRCA of

Anolis sensu stricto (Nicholson et al., 2012) and Norops during the middle Eocene (49 Mya,

HPD = 38-63 Mya). The MRCA of Amazonian species within Dactyloa, which corresponds to the split between A. dissimilis and the remaining Dactyloa taxa, dated back to the late Eocene (35

Mya, HPD = 25-47). The MRCA of Norops, which corresponds to the divergence of A. tandai from the other sampled Norops, dated back to the late Oligocene (24 Mya, HPD = 16-32 Mya).

The remaining diversification events between Amazonian anole species within both the Dactyloa and Norops clades happened during the Miocene (Fig. 1.6). For instance, within Dactyloa, the

MRCA of A. punctatus and A. phyllorhinus was dated around 12 Mya (HPD = 7-17 Mya), while the most recent common ancestor of the Norops anoles A. fuscoauratus and A. trachyderma was dated around 12 Mya (HPD = 7-17 Mya). Within Amazonian bush anoles, the dating analyses suggest that Polychrus liogaster and P. marmoratus diverged in the Pliocene (6 Mya, HPD = 3-

11 Mya).

18

Discussion

The phylogenetic affinities of Anolis dissimilis

Our analyses suggest that Anolis dissimilis is closely related to A. neblininus and A. calimae, two species distributed outside the Amazon basin and presenting highly disjunct distributions. While A. dissimilis occurs in southwestern Amazonia, A. calimae is restricted to the western Colombian Andes, and A. neblininus is a narrow endemic within the Guiana Shield.

Together, they composed a clade within Dactyloa that received low to maximum support in our analyses. Similar to our study, Castañeda and de Queiroz (2011) have previously recovered A. neblininus and A. calimae as sister taxa, yet independent of the five main clades within Dactyloa.

By contrast, Ayala et al. (1983) suggested that A. dissimilis and A. calimae are related to the punctatus species series, while Nicholson et al. (2012) proposed A. neblininus to be a member of the heterodermus series.

Due to limited genetic sampling of mainland Dactyloa, it is possible that the closest relatives of these three highly geographically separated taxa have not been included in our investigation. Morphological traits suggest unsampled mainland anole species that may be closely related to Anolis dissimilis, A. neblininus and A. calimae; for instance, Anolis deltae shares unique tail crests with A. dissimilis, while Anolis caquetae, Anolis santamartae and A. deltae exhibit a very large interparietal scale in contact with the semicircles, as seen in A. dissimilis (Williams, 1982). On the other hand, there has been no evident close relative of A. calimae (Ayala et al., 1983). To further verify whether the clade composed by A. dissimilis, A. calimae and A. neblininus represents a natural group, it will be key to continue improving the genetic sampling of mainland anoles.

19

Independent evolution of nasal appendages in anole lizards

We found that Anolis phyllorhinus is a member of the punctatus species series, and more specifically the sister species of A. punctatus, with which A. phyllorhinus is broadly sympatric.

Interestingly, with the exception of the proboscis and a red dewlap, A. phyllorhinus differs from

A. punctatus by only a few quantitative morphological traits. The identification of females, which lack both the proboscis and a developed dewlap, is indeed difficult (Rodrigues et al.,

2002).

Our data suggest that rostral appendages, found in Anolis phyllorhinus, A. proboscis and

A. laevis, have evolved at least twice within the anole radiation. Based on morphological comparisons, Yánez-Muñoz et al. (2010) also challenged the hypothesis of a close relationship between A. phyllorhinus and A. proboscis. Although no genetic data are currently available for A. proboscis or A. laevis, recent phylogenetic studies based on combined molecular and morphological evidence suggest that A. proboscis is closely related to the heterodermus species series (Castañeda and de Queiroz, 2013; Nicholson et al., 2012; Poe, 2004; Poe et al., 2012).

Morphological comparisons, in turn, led Williams (1979) to suggest that A. laevis is closely related to A. heterodermus (Williams, 1979). As a result, the laevis group, a name used to refer to the three anole species with a rostral proboscis (Williams, 1979), is not supported as a monophyletic group. Marked structural differences in the proboscises of A. phyllorhinus, A. proboscis and A. laevis (Williams, 1979) also support the view that these structures are not homologous (Yánez-Muñoz et al., 2010).

Genetic and phenotypic divergence between A. punctatus and A. philopunctatus

Our results indicate paraphyly of Anolis punctatus relative to A. philopunctatus, which is

20

nested among geographically close populations of A. punctatus from central Amazonia. We recovered short branches and low genetic distances between individuals with distinct dewlap patterns, including those with spotted (A. philopunctatus), orange (A. punctatus), and white (A. punctatus) dewlaps. Given the documented role of anole dewlaps in species recognition and presumably reproductive isolation (e.g., Losos, 1985; Macedonia et al., 1994; Macedonia and

Stamps, 1994; Sigmund, 1983), these results may indicate that A. philopunctatus is still in its initial stages of speciation. A vast number of studies have documented paraphyly among animal species that are phenotypically and often ecologically distinct (e.g., Brown and Twomey, 2009;

Johnson et al., 2005; McKay and Zink, 2010; Omland et al., 2000), including anoles (Thorpe and

Stenson, 2003). From the perspective of allele coalescence, the speciation process inherently starts with paraphyly in gene genealogies, and the probability of achieving reciprocal monophyly increases with time since population divergence (Knowles and Carstens, 2007).

Alternatively, our findings are also consistent with the hypothesis that the names Anolis philopunctatus and A. punctatus do not correspond, in fact, to distinct species (Avila-Pires,

1995). The degree of dewlap dissimilarity that disrupts species recognition and mating in Anolis lizards is unclear (Ng and Glor, 2011; Stapley et al., 2011). In addition to the dewlap, anoles rely on a range of visual signals, including body coloration and stereotyped head bobbing displays

(Jenssen and Gladson, 1984; Losos, 1985; Macedonia and Stamps, 1994; Muñoz et al., 2013).

Other studies have also found low genetic differentiation between distinct dewlap phenotypes

(D’Angiolella et al., 2011; Lambert et al., 2013; Ng and Glor, 2011; Stapley et al., 2011). Within some anole taxa, divergence in sexual signals has not prevented high levels of contemporary gene flow between populations (Muñoz et al., 2013), which might also be the case of A. punctatus and A. philopunctatus.

21

Because of the paraphyly of Anolis punctatus relative to A. philopunctatus, and given the low genetic distances among some of their sampled, we tentatively consider the name A. philopunctatus as a synonym of A. punctatus. However, spatially-dense sampling, along with a comprehensive analysis of morphological and genetic variation, is needed for a detailed taxonomic evaluation of A. philopunctatus, and of some highly genetically divergent populations within A. punctatus.

The age of Amazonian anoles

We found that most Amazonian taxa within both the Dactyloa and Norops clades of

Anolis diverged in the Miocene, but some diversification events were as old as the late Eocene and late Oligocene. Furthermore, Amazonian species of Polychrus diverged in the Pliocene.

Importantly, our divergence time estimates contradict the predictions of the Pleistocene Refugia

Hypothesis (Haffer, 1969; Vanzolini and Williams, 1970), one of the first models proposed for diversification in Amazonia. By examining distribution patterns and morphological variation within the Anolis chrysolepis species group, Vanzolini and Williams (1970) suggested that cycles of forest contraction and expansion during the Quaternary would have triggered repeated population isolation and divergence, promoting in situ allopatric speciation. However, our results disagree with the temporal framework implied in that model, since we recovered much older divergences between Amazonian anoles. Glor et al. (2001) also inferred pre-pleistocenic splits between the Amazonian anoles A. punctatus, A. chrysolepis, A. fuscoauratus, A. ortonii, A. scypheus and A. tandai. A scenario of Amazonian anole diversification preceding the Quaternary is consistent with molecular estimates emerging from a wide range of Amazonian organisms, because most crown-group ages date back to the Neogene (reviewed in Antonelli et al., 2010;

22

Hoorn et al., 2010).

Our inferred divergence times between Amazonian anole species are considerably younger than the most recent estimates (Nicholson et al., 2012), however. Nicholson et al. (2012) recovered the most recent common ancestor (MRCA) of Anolis chrysolepis, A. fuscoauratus, A. ortonii, and A. trachyderma at approximately 51.6 million years ago, more than twice our median estimate for the same node. They also found the MRCA of the Dactyloa and Norops clades to be around 95 million years old, nearly two times older than our estimate. Our analyses indicate that the sister lineages that ultimately originated Dactyloa and Norops (as currently recognized) did not diverge until the Eocene, in disagreement with the suggestion that the major modern clades of Anolis were present in the Caribbean region by the late Cretaceous. Although an evaluation of anole biogeography is beyond the scope of this work, our results suggest that existing hypotheses will benefit from the use of multiple nuclear markers, and of improved fossil data for node calibration.

Concluding remarks

Building on the Dactyloa clade molecular dataset of Castañeda and de Queiroz (2011), we provide new information about the phylogenetic placement of rare Amazonian anoles. The data reveal dewlap variation among genetically close forms within the punctatus species series, dispute the monophyly of the laevis group by supporting independent evolution of the anole proboscis, provide additional evidence for a sixth main clade within Dactyloa, and oppose the view that major modern anole clades were present in the Caribbean region by the late

Cretaceous. The findings lead us to new research questions about the , evolution and biogeography of anole lizards, which will build from denser sampling of continental species and

23

more extensive molecular datasets.

24

CHAPTER 2: Biogeographic links between southern Atlantic Forest and western South

America: rediscovery, redescription, and phylogenetic relationships of two rare montane

anole lizards from Brazil2

Introduction

Despite more than 250 years of biodiversity inventories, the rate of biological discoveries in the Neotropics remains high (Pimm et al., 2010; Scheffers et al., 2012), continuously transforming our understanding of regional biogeographic patterns and their underlying ecological and evolutionary processes (Angulo and Icochea, 2010). Biological discoveries often come from well-studied areas, as is the case of the Atlantic Rainforest, a biodiversity hotspot in eastern Brazil. In this region, recent expeditions have led to the description of several squamate and amphibian species (e.g., Rodrigues et al., 2007, 2009, 2013; Teixeira Jr. et al., 2012, 2013), and in some cases to the rediscovery of species that have remained undetected for decades (e.g.,

Pirani et al., 2010; Tonini et al., 2011; Zaher et al., 2005). Refining our knowledge about species ranges and phylogenetic relationships is key to improving inferences of historical biogeography and the drivers of diversification in the highly diverse and increasingly threatened Neotropical region (Angulo and Icochea, 2010).

In South America, our understanding of the historical processes that have shaped biodiversity patterns varies greatly across regions. Several investigations have found evidence of former connections and biotic exchange between the northern Atlantic Forest and eastern

Amazonia, which resulted in high taxonomic similarity among them (e.g., Batalha-Filho, et al.,

2 Published as: Prates I., Melo-Sampaio P.R., Drummond L.O., Teixeira Jr. M., Rodrigues M.T., Carnaval A.C. (2017). Biogeographic links between southern Atlantic Forest and western South America: rediscovery, re- description, and phylogenetic relationships of two rare montane anole lizards from Brazil. Molecular Phylogenetics and Evolution, 113: 49-58. 25

2013; Fouquet et al., 2012a,b; Prates et al., 2016a,b). By contrast, little is known about the historical processes that have shaped high levels of biodiversity in the cooler mountains that characterize the southern Atlantic Forest (Amaro et al., 2012). In this region, contemporary climate heterogeneity strongly correlates with lineage endemism (Carnaval et al., 2014), while phylogenetic patterns within bird, and rodent clades suggest historical connections and biotic exchange with the Andean Yungas and western Amazonia (e.g., Batalha-Filho, et al. 2013;

Castroviejo-Fisher et al., 2014, 2015; Percequillo et al., 2011). The disjunct distribution of some squamates further points to biogeographic links between the southern Atlantic Forest and western South American ecosystems, as in the well-known case of the of anguid lizard

Diploglossus fasciatus (Gray, 1831) (Vanzolini and Williams, 1970).

Ancient forest connections may explain the phylogenetic and taxonomic affinities of two rare endemic southern Atlantic Forest anole lizards, Anolis nasofrontalis Amaral, 1933 and

Anolis pseudotigrinus Amaral, 1933 (Fig. 2.1). So far, both species are represented by only a few specimens collected more than 40 years ago in two adjacent sites in the Brazilian state of Espírito

Santo. These two sympatric species are characterized by small to medium size, short limbs, lichenous coloration, and large smooth head scales. These traits have been interpreted as reminiscent, at least in part, of the Greater Antillean “twig anole” ecomorph, and therefore may provide evidence of adaptive convergence between mainland and Caribbean anoles (Poe et al.,

2015; Losos et al., 2012; Williams, 1976, 1992). Morphologically, A. nasofrontalis and A. pseudotigrinus contrast from the other three native Atlantic Forest anoles, Anolis fuscoauratus

D’Orbigny in Duméril and Bibron, 1837, Anolis ortonii Cope, 1868, and Anolis punctatus

Daudin, 1802, which occur predominantly in the northern Atlantic Forest lowlands (although expanding into a limited extent of southern Atlantic Forest). On the other hand, a twig anole-like

26

phenotype is also present in a number of South American species from the Andes and Guiana

Shield (e.g., Losos et al., 2012; Poe et al., 2015; Williams et al., 1976, 1996), which may be indicative of close phylogenetic relationships with A. nasofrontalis and A. pseudotigrinus.

However, because of the small number of collected specimens and lack of genetic samples, we know very little about the evolution and biogeographic relationships of these two rare Atlantic

Forest endemics.

Through targeted herpetological inventories in the Atlantic Forest, we recently obtained new samples of Anolis nasofrontalis and A. pseudotigrinus. We also identified specimens that have been deposited in zoological collections yet previously overlooked or misidentified. By combining new genetic data with published sequences of other species in the Dactyloa clade of

Anolis (Castaneda and de Queiroz, 2011; Poe et al., 2015; Prates et al., 2015), we investigate the phylogenetic relationships of A. nasofrontalis and A. pseudotigrinus and estimate divergence times from their closest relatives. Based on the morphological attributes of newly collected, previously collected, and type specimens, we provide a much-needed taxonomic redescription of

A. nasofrontalis and A. pseudotigrinus.

Material and Methods

Sampling of molecular data

Newly sampled Anolis nasofrontalis and A. pseudotigrinus were collected in the Reserva

Biológica Augusto Ruschi, state of Espírito Santo, coastal southeastern Brazil (-19.917, -40.552,

WGS1984). For molecular phylogenetic inference, we matched available genetic datasets (see below) and sequenced the mitochondrial gene NADH dehydrogenase subunit 2 (ND2) and the flanking tryptophan transfer RNA (tRNA-Trp) genes, following Jezkova et al. (2009), as well as

27

the nuclear recombination-activating gene 1 (RAG1), following Gartner et al. (2013).

Mitochondrial sequences obtained in Genbank often included four additional tRNAs flanking the

ND2 gene (tRNA-Ala, tRNA-Asx, tRNA-Cys, tRNA-Tyr) (Castañeda and de Queiroz, 2011), which were also used in our final alignments. Sequences were deposited in Genbank (accession numbers MF004396-9). For phylogenetic analyses, we combined our data with sequences generated by Ayala-Varela et al. (2014), Castañeda and de Queiroz (2011), Poe et al. (2015), and

Prates et al. (2015), totaling 58 sampled specimens from 56 Dactyloa clade Anolis species. We also included sequences of A. fuscoauratus, A. ortonii, Anolis tandai Avila-Pires, 1995, and

Anolis trachyderma Cope, 1876 (all within the Norops clade of Anolis) as outgroups.

Sequences were edited and aligned using Geneious Pro 6 (Biomatters, Auckland). We determined models of nucleotide evolution using Partition Finder 1.1.1 (Lanfear et al., 2012), implementing PhyML for likelihood estimation (Guindon and Gascuel, 2003) and the Bayesian information criterion for model selection (Sullivan and Joyce, 2005). Based on Partition Finder results, codon partitions were implemented for the protein-coding genes. Due to the small size of individual tRNA genes (~70 bp), which impaired proper estimation of substitution parameters in preliminary phylogenetic analyses, these regions were treated as a single partition.

Inferring phylogenetic relationships and divergence times

We performed simultaneous phylogenetic inference and divergence time estimation for the Dactyloa clade of Anolis under a Bayesian framework using BEAST 1.8.4 (Drummond et al.,

2012). An uncorrelated lognormal relaxed clock (Drummond et al., 2006) was implemented for each locus separately, as well as a birth-death process tree prior (Gernhard, 2008). To improve the estimation of nucleotide substitution rates, we set a lower prior bound of 0.0001 (instead of

28

the default, zero) to the corresponding rate parameters.

In the absence of Dactyloa anole fossils for time-calibration, we estimated divergence times through a secondary calibration strategy. For that, we relied on published divergence times among pleurodont iguanian lizards, including four Dactyloa species, which were estimated from sequences of five nuclear genes and three well-known fossils as calibration points (Prates et al.,

2015). Based on median divergence times and the associated 95% highest posterior density intervals (HPD) provided by Prates et al. (2015), we assigned a normally-distributed calibration prior to the node defining the most recent common ancestor (MRCA) of Anolis phyllorhinus

Myers and Carvalho, 1945, A. punctatus, and Anolis transversalis Duméril in Duméril and

Duméril, 1851, with a mean of 17.5 million years ago (mya) and a standard deviation of 3.5 mya.

Additionally, we set a normally-distributed calibration prior to the node defining the MRCA of those three species and Anolis dissimilis Williams, 1965, with a mean of 35.5 mya and standard deviation of 5.5 mya (this latter clade corresponds to the Continenteloa clade within Anolis proposed by Poe et al., 2015).

To parameterize the mean rate of the molecular clock (ucld.mean parameter in BEAST), we relied on substitution rates estimated by a historical demographic study of Anolis ortonii and

A. punctatus (Prates et al., 2016b), which included the loci used in the present investigation.

Based on mean mutation rates and the associated HPD from Prates et al. (2016b), we implemented a normally-distributed prior with mean of 1.40 x 10-2 substitutions per site per million year and standard deviation of 5 x 10-3 substitutions per site per million year for the mitochondrial locus (ND2 and flanking tRNAs), and a prior mean of 6.15 x 10-4, with standard deviation of 1.65 x 10-4, for the nuclear gene RAG1.

We performed three independent BEAST runs of 50 million generations each, with a

29

sampling frequency of 5,000 steps. After assessing convergence and stationarity of model parameters using Tracer 1.6 (Drummond et al., 2012), the three runs were combined in

LogCombiner 1.8.4 (with 10% of each run discarded as burn-in). A maximum clade credibility tree was then summarized in TreeAnnotator 1.8.4 (Drummond et al., 2012). Resulting topologies were visualized in FigTree 1.4 (available from http://tree.bio.ed.ac.uk/software/figtree).

Morphological analyses

Morphological measurements and scale characters follow Köhler (2014), with the exception of phalanx terminology, as we consider the concealed ungual phalanx as the first phalanx (following de Queiroz et al., 1998; ungual phalanx not considered in phalanx count by

Köhler, 2014). Measurements were made on preserved specimens with a digital caliper to the nearest 0.1 mm. Morphometric measurements are reported for adult specimens only. The scutellation characters of three specimens housed in the Naturhistorisches Museum Wien and one in the Zoologisches Museum Hamburg were scored based on high-resolution images of the head, body and limbs in dorsal, lateral and ventral views. However, no morphometric measurements were made based on photographs. Color in life was extracted from field notes and photographs of recently collected specimens. To support biological inventories in the Atlantic

Forest, we provide comparisons of Anolis nasofrontalis and A. pseudotigrinus with the other native Atlantic Forest anole species (A. fuscoauratus, A. ortonii, and A. punctatus) based on

Ávila-Pires (1995).

Results

Phylogenetic relationships and divergence times

30

Phylogenetic analyses recovered six main clades within Dactyloa (Fig. 2.2), five of which were previously referred to as the aequatorialis, heterodermus, latifrons, punctatus, and roquet species series (Castañeda and de Queiroz, 2013; Poe et al., 2015; Prates et al., 2015). A sixth clade clustered the newly sampled Anolis nasofrontalis and A. pseudotigrinus along with A. dissimilis, Anolis calimae Ayala et al., 1983, Anolis neblininus (Myers, Williams and

McDiarmid, 1993), and two undescribed Andean species (Anolis sp. R and Anolis sp. W from

Poe et al., 2015). Each of these six major Dactyloa clades received maximum support, yet the relationships between them were, generally, poorly supported. The only exception is the well- supported sister relationship between the aequatorialis and latifrons species series (Fig. 2.2).

The Atlantic Forest species Anolis nasofrontalis and A. pseudotigrinus were recovered as sister taxa with high support (Fig. 2.2). The MRCA of these two anoles is the sister of A. dissimilis, a western Amazonian anole from Brazil and Peru. Results indicate that A. pseudotigrinus and A. nasofrontalis are not close relatives of A. punctatus, the only other

Atlantic Forest lizard within the Dactyloa clade of Anolis. Divergence time estimates indicate that A. pseudotigrinus and A. nasofrontalis share a MRCA at 9.92 mya (HPD = 6.55–13.49 mya), while the MRCA of these two species diverged from its sister A. dissimilis around 11.49 mya (HPD = 8–15.36 mya). The MRCA of A. nasofrontalis, A. pseudotigrinus, A. dissimilis, A. calimae, and A. neblininus was dated as being around 27.14 my old (median value; HPD =

19.96-35.29 mya).

Within the Dactyloa clade, divergences between most sampled species predate the

Quaternary (i.e., are older than ~2.6 mya), with most such events dating to the Miocene (~5.5-23 mya; Fig. 2.2). All six major clades within Dactyloa began to diversify during the early Miocene and late Oligocene, between 19.33 and 27.14 mya. We found the MRCA of all sampled Dactyloa

31

to be 32.58 my old (HPD = 24.59–41.27 mya), while the MRCA of the Dactyloa and Norops clades of Anolis (the latter represented by A. fuscoauratus, A. ortonii, A. tandai, and A. trachyderma) was recovered as around 49.31 my old (HPD = 35.72–64.36 mya). Despite the use of different sets of calibrated nodes and the incorporation of a mitochondrial marker in the present study, age estimates for these deeper nodes are very similar to those of Prates et al.

(2015).

Taxonomic accounts

Taxon redescription: Anolis nasofrontalis Amaral, 1933

Fig. 2.1a, b; Fig. 2.2; Fig. 2.3a, b

Anolis nasofrontalis – Myers and Carvalho, 1945: 6, 9, 14; Williams and Vanzolini,

1980: 99, 103–106; Williams 1992: 11, 12, 15, 16, 22; Castañeda and de Queiroz, 2013: 350,

375, 379, 380; Poe et al., 2015: 640, 641, 646, 650.

Dactyloa nasofrontalis – Nicholson et al., 2012: 83, 95.

Holotype: MZUSP 440, adult female from Espírito Santo (ES), Brazil, collected by

Ernesto Garbe in 1906.

Paratype: MZUSP 440.A, subadult male (the allotype). Same collection data as the holotype. The poor preservation state of this specimen hindered its inclusion in the characterization of Anolis nasofrontalis.

Additional specimens examined: ZMH R0411, a female from Santa Leopoldina, ES, collected by W. Schlüter on December 1902; NHMW 12742, NHMW 25201.1, and NHMW

32

25201.2, females from Santa Leopoldina, ES, from the collection of F. Werner (who worked in the Naturhistorisches Museum Wien from 1919 to 1939); MNRJ 1625, an adult female from

Santa Teresa, ES, collected by A. Ruschi on January 1940; LOD 1383, an adult female collected at Reserva Biológica Augusto Ruschi, Santa Teresa, ES, by L. O. Drummond and P. R. Melo-

Sampaio on April 2016.

Characterization: Snout-vent length 51.4 to 55.1 mm; head length 12.8 to 13.7 mm, width 8.3 to 8.4 mm; snout length 7 to 8 mm; ear opening 0.9 to 1.3 mm horizontally, 0.7 vertically; interparietal scale width 1.4 mm, length 2.2 mm; postcloacal scales not visible in male specimen, likely due to poor preservation; tail length 76.1 to 85 mm; tail round in cross section; fourth toe of adpressed hind limb not reaching posterior insertion of arm; shank length 9.1 to

10.2 mm; fourth finger of extended forelimb reaching about anterior border of eye; width of dilated subdigital pad 0.9 mm, of non-dilated subdigital pad 0.2 mm.

Dorsal head scales smooth and juxtaposed (Fig. 2.3a), ventral ones smooth and granular

(Fig. 2.3b); deep prefrontal depression; rostral slightly overlaps mental; four to six postrostrals; single prenasal in contact with rostral; circumnasal in contact with rostral and first supralabial; four to five internasals; six to eight supralabials and five to eight infralabials to the level of center of eye; three to seven postmentals, the two outer ones greatly enlarged; two to five enlarged sublabials in contact with infralabials; canthal ridge indistinct to well-defined; eight to

17 smooth loreals in two to four rows at level of second canthal; five to seven scales between first canthals, four to seven between second canthals; three to four smooth suboculars in a single row, in contact with supralabials; one to three slightly elongated smooth superciliaries above anterior portion of eye; two to three greatly enlarged smooth supraoculars; one incomplete row of circumorbitals separating supraoculars from supraorbital semicircles; semicircles broadly in

33

contact; parietal depression shallow to moderate; interparietal distinctly enlarged, surrounded by large to moderate-sized scales; semicircles in contact with interparietal, or separated by one scale; scales anterior and posterior to ear opening subequal, slightly larger anteriorly.

Dewlap small in males and females; anterior insertion of dewlap at level of posterior border of eye, posterior insertion at level of anterior insertion of arm; about five to seven rows of gorgetal-sternal dewlap scales, about twice as large as ventrals, separated by naked skin.

Dorsal body scales smooth and juxtaposed, mostly hexagonal; no enlarged rows of dorsals; 78-89 dorsals longitudinally between level of axilla and groin; lateral scales smooth, granular, rounded; ventrals smooth, imbricate, with rounded posterior margins, about twice the size of dorsals; 66–73 ventrals longitudinally between level of axilla and groin; 95–104 scales around mid-body; axillary depression or pocket absent; proximal portion of tail (about 1/5 of tail length) with smooth, mostly hexagonal scales, distal portion with keeled, imbricate, mostly hexagonal scales.

Supradigital scales smooth; 27–32 subdigital lamellae under phalanges III-V of fourth toe, six to seven scales under second phalanx; ventral surface of second phalanx raised relative to that of proximal phalanges; subdigital lamellae of toepad projecting distally under subdigital scales of second phalanx.

Color in life: Dorsal surface of body, tail and limbs lichenous with light-gray background

(Fig. 2.1a,b); diagonal lines composed of dark-brown spots on the first half of body, irregularly- distributed dark reticulations and scattered yellow spots on the second half; dark and well- marked interorbital line on the dorsal surface of head; two dark lines radiating from each eye, the first one extending posteriorly halfway to tympanum, the second anteriorly through canthus rostralis; iris brown; tail with nine dark-brown bands; ventral surface of head, body, tail and

34

limbs light-gray with scattered dark-brown spots; throat lining black (Fig. 2.1a); tongue tan; female dewlap skin light-orange, faded to white posteriorly; dewlap scales white; based on

Amaral (1933), male dewlap skin pink.

Species comparisons (traits of taxa compared to Anolis nasofrontalis in parentheses): A. nasofrontalis can be distinguished from A. fuscoauratus, A. ortonii, and A. punctatus by 95–104 scales around mid-body (A. fuscoauratus: 124–157; A. ortonii: 123–180; A. punctatus: 132–167), smooth dorsals (weakly keeled), and smooth snout scales between nostrils (keeled). Anolis nasofrontalis further differs from A. fuscoauratus and A. punctatus by a lichenous dorsal body coloration (A. fuscoauratus: brown or gray; A. punctatus: green, often with white spots), supraorbital semicircles in contact (separated by granular scales) and well-developed dewlap in females (vestigial). Anolis nasofrontalis can be distinguished from A. pseudotigrinus by about five to seven rows of gorgetal-sternal dewlap scales (12–14), 78–89 dorsals (87–98) and 66–73 ventrals (79–90) between axilla and groin, 95–104 scales around mid-body (106–118), and a small light orange dewlap in females (large white dewlap). Anolis nasofrontalis can be further distinguished from A. pseudotigrinus and A. punctatus by a blunt snout (pointed).

Distribution and natural history: Anolis nasofrontalis is presently confirmed to occur in the Reserva Biológica Augusto Ruschi in Santa Teresa, ES, at an altitude of 745 m above sea level (asl). The local vegetation is characterized by dense broad-leaf montane Atlantic rainforest.

The recently-collected specimen was found sleeping at night on a narrow branch at a height of about 3 m. This species is sympatric with A. fuscoauratus, A. pseudotrigrinus, and A. punctatus.

Four additional A. nasofrontalis (housed in Vienna and Hamburg) were identified as collected in

Santa Leopoldina, ES, a lowland site (~85 m asl) situated about 18 km away from Santa Teresa; yet, this name was previously used to designate a municipality of large territorial extent that was

35

dismembered into smaller districts, including the Comuna de Santa Teresa (currently the municipality of Santa Teresa). We failed to find A. nasofrontalis in Santa Leopoldina; its presence in this area is yet to be confirmed.

Taxon redescription: Anolis pseudotigrinus Amaral, 1933

Fig. 2.1c, d; Fig. 2.2; Fig. 2.3c, d

Anolis pseudotigrinus – Myers and Carvalho, 1945: 6; Williams and Vanzolini, 1980: 99,

103–106; Williams 1992: 11, 12, 15, 16, 22; Castañeda and de Queiroz, 2013: 350, 375, 379,

380; Poe et al., 2015: 640, 641, 646, 650.

Dactyloa pseudotigrina – Nicholson et al., 2012: 83, 96.

Holotype: MZUSP 721.B, adult female from the Rio Doce region, ES, collected by

Ernesto Garbe in 1906.

Additional specimens examined: MZUSP 36718, an adult male collected in Santa Teresa,

ES, by J. F. Jackson on February 1974; MBML 327, an adult female, MBML 266 and 536, two adult males, and MBML 554, a subadult male, all collected in the park of the Museu de Biologia

Professor Melo Leitão in Santa Teresa, by M. G. Hoffmann in 1997-8; MTR 34789 and 34790, two adult females collected in the Reserva Biológica Augusto Ruschi, ES, by M. Teixeira Jr., M.

T. Rodrigues, F. Dal Vechio, R. S. Recoder and R. Damasceno on December 2014; LOD 1237, a subadult female from Reserva Biológica Augusto Ruschi, collected by L. O. Drummond and P.

R. Melo-Sampaio on February 2016.

Characterization: Snout-vent length 56.4 to 65.8 mm; head length 15.4 to 17.2 mm,

36

width 7.2 to 8.1 mm; snout length 8.4 to 9.8 mm; ear opening 0.7 to 1.0 mm horizontally, 0.8 to

1 mm vertically; interparietal scale width 1.9 to 2.4 mm, length 2.3 to 2.7 mm; postcloacal scale width in males 0.6 to 1.3 mm; tail length 101 to 112 mm; tail round in cross section; fourth toe of adpressed limb reaching posterior insertion of arm; shank length 10.2 to 11.1 mm; fourth finger of extended forelimb reaching about anterior border of eye; width of dilated subdigital pad 1.2 to

1.5 mm, of non-dilated subdigital pad 0.3 to 0.4 mm.

Dorsal head scales smooth and juxtaposed (Fig. 2.3c), ventral ones smooth and granular

(Fig. 2.3d); deep prefrontal depression; rostral strongly overlaps mental; four to seven postrostrals; one to two prenasals, lower one in contact with rostral; circumnasal in contact with rostral and first supralabial; four to six internasals; seven to nine supralabials and infralabials to the level of center of eye; four to five postmentals, the two outer ones greatly enlarged; two to four enlarged sublabials in contact with infralabials; canthal ridge indistinct; 16 to 24 smooth loreals in three to five rows at level of second canthal; six to seven scales between first canthals, five to seven between second canthals; four to five smooth suboculars in a single row, in contact with supralabials; one to two smooth to keeled elongate superciliaries above anterior portion of eye; two to five greatly enlarged smooth supraoculars; one incomplete row of circumorbitals separating supraoculars from supraorbital semicircles; semicircles broadly in contact; parietal depression ill-defined; interparietal distinctly enlarged, surrounded by large to moderate-sized scales; semicircles in contact with interparietal, or separated by one scale; scales anterior and posterior to ear opening subequal to slightly larger anteriorly.

Dewlap large in males and females; anterior insertion of dewlap at level of anterior border of eye, posterior insertion at level of posterior insertion of arm (females) or reaching about 1/3 to 1/2 of length between axilla and groin (males); about 12–14 rows of gorgetal-sternal

37

dewlap scales, about twice as large as ventrals, separated by naked skin.

Dorsal body scales smooth, juxtaposed, mostly hexagonal; no enlarged rows of dorsals;

87-98 dorsals longitudinally between level of axilla and groin; lateral scales smooth, granular, rounded; ventrals smooth, imbricate, with rounded posterior margins, about twice the size of dorsals; 79–90 ventrals longitudinally between level of axilla and groin; about 106–118 scales around mid-body; axillary depression or pocket absent; proximal portion of tail (about 1/4 to 1/5 of tail length) with smooth, mostly hexagonal scales, distal portion with keeled, imbricate, mostly hexagonal scales; in males, two moderately enlarged postcloacals.

Supradigital scales smooth; 26–30 subdigital lamellae under phalanges III-V of fourth toe, six to seven scales under second phalanx; ventral surface of second phalanx raised relative to that of proximal phalanges; subdigital lamellae of toepad projecting distally under subdigital scales of second phalanx.

Color in life: Dorsal surface of body, tail and limbs lichenous with light-green to olive- green background (Fig. 2.1c,d); diagonal lines composed of dark-green or bluish spots on first half of body, irregularly-distributed green reticulations and green spots on second half; dorsal surface of head green; large brown nuchal spot; no dark interorbital line; iris brown; four large mid-dorsal brown spots on body, first one at the level of anterior limb insertion, fourth one at the level of posterior limb insertion; tail with four to six brown spots or bands; ventral surface of head, body, tail, and limbs light-gray with scattered dark-brown spots; throat lining black; tongue pink; female dewlap skin white; dewlap scales white; male dewlap coloration unknown.

Species comparisons (traits of taxa compared to Anolis pseudotigrinus in parentheses):

Anolis pseudotigrinus can be distinguished from A. fuscoauratus, A. ortonii, and A. punctatus by a lichenous dorsal body coloration with green background (A. fuscoauratus: brown or gray; A.

38

ortonii: marmorated with brown, gray and black; A. punctatus: green, often with white spots),

106–118 scales around mid-body (A. fuscoauratus: 124–157; A. ortonii: 123–180; A. punctatus:

132–167), smooth dorsals (weakly keeled), and smooth snout scales between nostrils (keeled).

Anolis pseudotigrinus further differs from A. fuscoauratus and A. punctatus by supraorbital semicircles in contact (separated by granular scales) and well-developed dewlap in females

(vestigial). Anolis pseudotigrinus differs from A. nasofrontalis by about 12–14 rows of gorgetal- sternal dewlap scales (five to seven), 87–98 dorsals (78–89) and 79–90 ventrals (66–73) between axilla and groin, 106–118 scales around mid-body (95–104), and large white dewlap in females

(small orange dewlap). Anolis pseudotigrinus can be further distinguished from A. fuscoauratus,

A. nasofrontalis, and A. ortonii by a pointed snout (blunt).

Distribution and natural history: Anolis pseudotigrinus is known only from the Reserva

Biológica Augusto Ruschi (~745 m asl) and from the forest reserve of the Museu de Biologia

Professor Melo Leitão (~700 m asl), both in Santa Teresa, ES. It is sympatric with A. fuscoauratus, A. nasofrontalis and A. punctatus. The local vegetation is characterized by dense broad-leaf montane Atlantic rainforest. Specimens were collected during the night, sleeping on narrow branches at a height of 2–3 m, one of them on a bamboo cluster near forest edge.

Discussion

Phylogenetic relationships and taxonomic implications

Historical relationships among South American anoles have been controversial, yet much improved by genetic data, which have revealed unexpected phylogenetic and biogeographic associations (e.g., Castañeda and de Queiroz, 2011; Nicholson, 2002; Nicholson et al., 2005;

Poe, 2004; Prates et al., 2015). Traditionally, Anolis nasofrontalis and A. pseudotigrinus have

39

been assigned to the tigrinus species group, which also included species from the Colombian and

Venezuelan Andes (Williams, 1976, 1992). Because molecular data recovered the tigrinus group as partially nested within the punctatus species series, that group designation is no longer used

(Castañeda and de Queiroz, 2011). Subsequently, A. nasofrontalis and A. pseudotigrinus were tentatively assigned to the punctatus species group (Nicholson et al., 2012), and, more recently, considered incertae sedis within Dactyloa due to poor resolution in phylogenetic analyses that incorporated both genetic and morphological data (Castaneda and de Queiroz, 2013; Poe et al.,

2015). Based on new DNA sequence data, we were able to refine the phylogenetic placement of these two poorly-known Brazilian species.

New genetic data from Anolis nasofrontalis and A. pseudotigrinus have implications for

Anolis taxonomy. Similar to our study, previous molecular phylogenetic analyses found A. neblininus and A. calimae (Castaneda and de Queiroz, 2011), or these two species along with A. dissimilis (Prates et al., 2015), to compose a clade within Dactyloa. This clade fell outside of the aequatorialis, heterodermus, latifrons, punctatus, and roquet species series (Castaneda and de

Queiroz, 2013; Poe et al., 2015), suggesting a sixth major Dactyloa clade. In agreement with this scenario, our analyses found A. nasofrontalis, A. pseudotigrinus, A. calimae, A. dissimilis, A. neblininus, and two undescribed Andean species (Anolis sp. R and Anolis sp. W from Poe et al.,

2015) to compose an exclusive clade. In the face of these findings, and for practical reasons, we define a sixth species series within the Dactyloa clade of Anolis: the neblininus species series. It is named after the neblininus group of Williams et al. (1996), composed by the Guiana Shield endemics A. neblininus and Anolis carlostoddi (Williams, Pradero and Gorzula, 1996) and formerly allocated to the genus Phenacosaurus Lazell, 1969 (a genus later synonymized with

Anolis; Poe, 1998). The inferred composition of the neblininus species series includes A.

40

calimae, A. dissimilis, A. nasofrontalis, A. neblininus, and A. pseudotigrinus based on our phylogenetic results, as well as A. carlostoddi in accordance with Williams et al. (1996).

Further sampling of South American anole taxa may reveal additional members of the neblininus species series, perhaps with implications for the phylogenetic affinities of Anolis nasofrontalis and A. pseudotigrinus. For instance, these two species share with the central

Andean Anolis williamsmittermeierorum Poe and Yañez-Miranda, 2007 a black throat lining, a seemingly unusual trait (Poe and Yañez-Miranda, 2007) that may be indicative of close phylogenetic relationships. In a study incorporating both molecular and morphological characters

(Poe et al., 2015), A. williamsmittermeierorum and its relative Anolis peruensis Poe, Latella,

Ayala-Varela, Yañez-Miranda and Torres-Carvajal, 2015 were found to be closely related to A. neblininus, a close relative of A. nasofrontalis and A. pseudotigrinus in our analysis. Poe et al.

(2015) also found A. peruensis and A. williamsmittermeierorum to be closely related to A. carlostoddi, a member of Williams' former neblininus group, and to Anolis bellipeniculus (Myers and Donelly, 1996), another tepui (table mountain) endemic. Based on these findings, we expect that the neblininus series will likely expand to include Andean and Guiana Shield species for which genetic data is not yet available, as well as still unnamed anoles such as Anolis sp. R and

Anolis sp. W (Poe et al., 2015).

Along with other members of the neblininus species series, Anolis nasofrontalis and A. pseudotigrinus exhibit morphological attributes that have been associated with the Greater

Antillean “twig anole” ecomorph, such as short limbs and cryptic coloration (Williams, 1976,

1992). This twig anole-like morphology has been also attributed to Andean anoles currently assigned to the punctatus (e.g., Anolis tigrinus Peters, 1863) and heterodermus series (e.g.,

Anolis euskalerriari (Barros, Williams and Viloria, 1996), Anolis orcesi (Lazell, 1969), and

41

Anolis proboscis Peters and Orces, 1956) (Losos et al., 2012; Poe et al., 2015; Williams et al.,

1996). Phylogenetic relationships within Dactyloa support the view of recurrent evolution of a twig anole-like phenotype in South America, perhaps as a result of adaptive convergence; alternatively, this pattern may reflect the conservation of a phenotype ancestral for the Dactyloa clade (Castañeda and de Queiroz, 2011; Poe et al., 2015). In the former case, an apparent association with South American mountains is intriguing. Nevertheless, natural history data are still needed to assess whether mainland anoles exhibit the typical ecological and behavioral traits that characterize Caribbean anole ecomorphs – in the case of twig anoles, active foraging, slow movements, infrequent running or jumping, and preference for narrow perching surfaces (Losos,

2009). Although scarce, the available data support that at least some twig anole-like mainland species behave as typical Caribbean twig anoles, as is the case of Anolis proboscis (Losos et al.,

2012).

Biogeographic connections between South American highlands

Species in the neblininus series have been sampled in localities separated by large geographic distances (Fig. 2.4). While Anolis nasofrontalis and A. pseudotigrinus are restricted to the southern montane Atlantic Forest, A. dissimilis is known from a few sites in southwestern

Amazonia adjacent to the Andean foothills, A. calimae occurs on the western cordillera of the

Colombian Andes, and A. neblininus is a narrow tepui endemic in the Guiana Shield, close to the

Brazil-Venezuela border. Such purported relationships between Andean, Guiana Shield, and southern Atlantic Forest taxa point to an assortment of narrowly-distributed mainland anoles that are associated with mid elevations, or their adjacent habitats in the case of A. dissimilis (Fig.

2.4). A similar pattern has been found by a number of phylogenetic studies of South American

42

frog clades (e.g., Castroviejo-Fisher et al., 2014, 2015; Fouquet et al., 2012a,b; Padial et al.,

2015).

Close phylogenetic relationships between these highly disjunct and narrowly distributed lizards seem to indicate a complex biogeographic history involving former patches of suitable habitat between regions, followed by habitat retraction and extinction in the intervening areas. In the case of Anolis nasofrontalis and A. pseudotigrinus, past forest corridors may explain a close relationship with the western Amazonian A. dissimilis. Atlantic and Amazonian rainforests are presently separated by the open savannas and scrublands of the Caatinga, Cerrado and Chaco domains. However, geochemical records suggest that pulses of increased precipitation and wet forest expansion during the Quaternary have favored intermittent connections between western

Amazonia and the southern Atlantic Forest (through present-day southwestern Brazil), as well as between eastern Amazonia and the northern Atlantic Forest (through northeastern Brazil) (Cheng et al., 2013). Genetic data indicate that a northern forest connection has led to colonization of the

Atlantic Forest from eastern Amazonia by A. ortonii and A. punctatus, as well as by the bush anole Polychrus marmoratus (Linnaeus, 1758), in the mid-Pleistocene (Prates et al., 2016a). By contrast, our findings indicate a much older, Miocene divergence between A. dissimilis and the ancestor of A. nasofrontalis and A. pseudotigrinus. Similar to our results, Miocene divergences between disjunct sister species and clades that occur in southern Atlantic Forest and western

Amazonia have been reported for a number of bird and frog taxa (e.g., Batalha-Filho et al., 2013;

Gehara et al., 2014). These old divergences suggest that intermittent southern forest connections in South America may largely predate the time frame encompassed by most available paleoenvironmental data (i.e., the Pleistocene; Cheng et al., 2013).

These inferred biogeographic connections in South America may have also been

43

influenced by major landscape shifts as a result of Andean orogeny during the mid to late

Miocene. The timing of divergence between Anolis dissimilis and the ancestor of A. nasofrontalis and A. pseudotigrinus (~11.5 mya) overlaps with that of the Chapare buttress (Fig.

2.4), which was a land bridge that connected the central Andes to the western edge of the

Brazilian Shield, around present-day (Lundberg et al., 1998). This bridge may have acted as a path for west-east movement of terrestrial organisms isolated by the paleo-Amazonas (to the north) and Paraná (to the south) river systems, and later by the Pebasian and Paranan inland seas

(Wesselingh and Salo, 2006). Additionally, wet environments associated with the Paraná system may have acted as a corridor between this region and Brazil's southeast (Batalha-Filho et al.,

2013), favoring biotic exchange between the southern Atlantic Forest with the Andean Yungas and adjacent western Amazonia. Although unconfirmed, the historical records of A. nasofrontalis in a lowland site (Santa Leopoldina) may indicate some tolerance to lowland conditions in this group of anoles, in agreement with the presence of the closely-related A. dissimilis in lowland western Amazonia. Such tolerance may have played a role in the colonization of the Atlantic

Forest from western South America, through occupation of the intervening lowlands.

Interestingly, geological evidence also supports connections between the northern Andes and the Guiana Shield during the mid-Miocene through the Vaupés Arch (Lundberg et al., 1998), which is consistent with our divergence time estimates between Andean and Guianan highland anoles from the neblininus species series.

Concluding remarks

Building upon the molecular scaffolding first provided by Castañeda and de Queiroz

(2011) for the Dactyloa clade of Anolis, our knowledge about this group's diversity and evolution

44

has increased rapidly. Based on new biological inventories and genetic data, we have uncovered the phylogenetic relationships and provided a taxonomic redescription of A. nasofrontalis and A. pseudotigrinus, two rare endemic Atlantic Forest anoles. We find evidence of a Miocene rainforest corridor between southeastern Brazilian forests and western Amazonia, and biogeographic associations with similarly narrowly-distributed species from the Andes and the

Guiana Shield. Based on the phylogenetic results, we define the neblininus species series, a sixth major clade within the Dactyloa radiation of Anolis.

It has become increasingly clear that broader sampling of genetic variation is essential in advancing studies of mainland anole taxonomy and evolution. This significant challenge also provides exciting opportunities for complementary sampling efforts, exchange of information, and new collaborations between research groups working in different South American countries.

45

CHAPTER 3: A mid-Pleistocene rainforest corridor enabled synchronous invasions of the

Atlantic Forest by Amazonian anole lizards3

Introduction

Shifts in the geographic distribution of habitats over time can promote dispersal and vicariance, thereby influencing large-scale biogeographic patterns and ecological processes. For instance, topographic discontinuities (e.g. mountains, rivers) and environmental barriers (e.g. dry corridors separating wet ecosystems) can prevent the homogenization of species pools while favoring genetic divergence between closely related disjunct taxa (Graham et al., 2004).

Whenever those barriers are disrupted, organisms may have an opportunity to colonize adjacent suitable areas and hence encounter and interact with new species and environments (Petren et al.,

2005). The ecological opportunities promoted by these habitat shifts may lead to evolutionary radiations (Hughes and Eastwood, 2006).

The establishment of suitable habitat corridors across previously disjunct, ecologically similar regions has been widely linked to climate change over time. For example, the allopatric distribution of closely related mountaintop endemics has been tied to downhill dispersal during former cold periods, followed by uphill movement in subsequent warmer times, allowing species to reach nearby peaks (Kozak and Wiens, 2006). Similarly, climate-mediated sea level fluctuation has been invoked in cross-island divergences (Brown et al., 2013). These cases support the claim that climatic variation plays a major role in the establishment of spatial patterns of biodiversity not only by constraining species ranges (e.g., Carnaval et al., 2014), but also by enabling species exchange between biologically distinct geographic regions.

3 Published as: Prates I., Rivera D., Rodrigues M.T., Carnaval, A.C. (2016). A mid-Pleistocene rainforest corridor enabled synchronous invasions of the Atlantic Forest by Amazonian anole lizards. Molecular Ecology, 25: 5174– 5186. 46

Climate-mediated dispersal is thought to have played a key role in the assembly of tropical biotas, including those of Amazonia and the Atlantic Forest in South America, two of the most diverse ecosystems on Earth. These rainforest blocks are currently separated by the

Caatinga (xeric scrublands) and the Cerrado (savannas) domains, which prevent dispersal of rainforest-associated organisms. However, there is substantial evidence that Amazonian and

Atlantic Forest biotas have been linked through time. Ubiquitous biotic exchange in the past is supported by present-day floristic similarity (Bigarella et al., 1975; Santos et al., 2007) and disjunct distribution patterns of many animal and plant species (e.g., Fiaschi and Pirani, 2009;

Gehara et al., 2014; Ribeiro-Júnior, 2015; Rocha et al., 2015). Sister relationships between species and clades restricted to either Amazonia or the Atlantic Forest indicate that in situ diversification following dispersal has contributed to the high levels of endemism that characterize both systems (e.g., Batalha-Filho et al., 2013; Costa, 2003; Fouquet et al., 2012a, b).

However, limited data are available on the timing, frequency, and magnitude of these former connections. Did they happen through narrow corridors of suitable habitat, or was former climate change enough to enable major fusion of the forest domains that are now physically isolated?

Paleoenvironmental data shed light on the climatic drivers of former biotic exchange between major South American rainforests. Oxygen isotope records (Cheng et al., 2013), speleothem and travertine deposition patterns (Auler et al., 2004), and paleoclimatic simulations

(Sobral-Souza et al., 2015) suggest that climatic variation through time has favored pulses of rainforest expansion. These data support a scenario of intermittent periods of increased precipitation that date back at least 900 ky (Auler et al., 2004) and match the oscillations of the precession component of Earth’s orbital cycles (every ~20 ky; Cheng et al., 2013). It has been hypothesized that phases of increased humidity during the Quaternary may have enabled the

47

establishment of forest corridors between eastern Amazonia and the northern Atlantic Forest

(through present-day northeastern Brazil), as well as between southwestern Amazonia and the southern Atlantic Forest (through southwestern Brazil). These two routes are consistent with the distribution patterns of bird species that occur disjunctively in both domains (Cheng et al., 2013), as well as with spatial patterns of phylogenetic structure observed among other vertebrate taxa

(e.g., Costa, 2003; Fouquet et al., 2012a,b; Gehara et al., 2014). However, most divergence time estimates indicate pre-Quaternary splits between Amazonian and Atlantic Forest lineages (i.e., older than ~2.6 mya), which precedes the time frame of the paleoenvironmental data available by millions of years (e.g., Batalha-Filho et al., 2013; Costa, 2003; Fouquet et al., 2012a,b; Gehara et al., 2014; Rodrigues et al., 2014). Here, we present an investigation of intra-specific genetic structure in species that occur in both forests - whose populations presumably diverged more recently - to more effectively address the impact of Late Quaternary climate change on the connectivity of South American rainforests.

To investigate the magnitude of former connections between Amazonia and the Atlantic

Forest, we capitalize on the fact that demographic syndromes such as population expansion and bottlenecks leave contrasting genetic signatures in the local biota (Beaumont et al., 2002;

Carnaval et al., 2009; Fagundes et al., 2007), and use current patterns of genetic diversity to test alternative biogeographic and demographic hypotheses. Specifically, if short-term or spatially restricted forest corridors enabled movement between Amazonia and the Atlantic Forest, we expect to detect a signal of colonization by a small subset of individuals from the source population into the colonized region. In this case, a population bottleneck would have occurred along with dispersal into the new area, followed by population expansion. Alternatively, long- lasting or extensive connections might have favored high levels of gene flow between Amazonia

48

and the Atlantic Forest. Subsequent isolation of forest blocks could have led to vicariance and divergence, with no population bottlenecks. Finally, if forest connections happened cyclically - as suggested by the documented climatic fluctuations (Cheng et al., 2013) - one also may expect that populations in different forest blocks have experienced some level of gene flow after divergence.

To assess these alternative historical hypotheses, we utilize extensive geographic sampling of three distantly-related lizard species (divergence > 49 mya; Prates et al., 2015): the true anoles Anolis punctatus and Anolis ortonii (Dactyloidae), and the bush anole Polychrus marmoratus (Polychrotidae). These species are similar in their arboreal habits and association to forests, but differ in their tolerance to transitional habitats: P. marmoratus explores forest edges, while the two Anolis are associated with closed-canopy forests (Kawashita-Ribeiro and Ávila,

2008; Ribeiro-Júnior, 2015; Vitt et al., 2003). The ancestral ranges of these three species have been traced back to Amazonia, but the Amazonian regions that served as sources of expansion remain unclear (Prates et al., 2016a). To allow proper biogeographic reconstruction, we perform targeted sampling in areas that have likely been connected by forest corridors, as suggested by the available paleoclimatic (Cheng et al., 2013; Sobral-Souza et al., 2015) and phylogenetic

(Batalha-Filho et al., 2013; Costa, 2003; Gehara et al., 2014) data. To test between alternative historical scenarios, we use coalescent simulations, Approximate Bayesian Computation

(Beaumont, 2010; Csillery et al., 2010), and empirical sequence data at six unlinked loci from

159 sampled specimens. Employing a hierarchical Approximate Bayesian Computation approach

(Hickerson et al., 2007), we then test whether colonization of the Atlantic Forest happened synchronously across the three species.

49

Material and methods

Molecular sampling

We generated DNA sequences for 102 specimens of Anolis punctatus (Dactyloidae), 34

A. ortonii, and 23 Polychrus marmoratus (Polychrotidae) collected in Brazil (list of voucher numbers and locality information in Appendix 3.1). Our sampling targeted localities in the

Atlantic Forest as well as in eastern and southwestern Amazonia, which were presumably connected to the Atlantic Forest through forest corridors (Cheng et al., 2013). As outgroups, we sampled two individuals of each of the following species: Anolis dissimilis, Anolis fuscoauratus,

Anolis phyllorhinus, Anolis scypheus, Anolis tandai, Anolis trachyderma, Anolis transversalis, and Polychrus liogaster. The mitochondrial gene NADH dehydrogenase subunit 2 (ND2) and the flanking tryptophan transfer RNA (tRNA-Trp) gene were sequenced as per Jezkova et al. (2009).

Additionally, five nuclear genes were sequenced: the recombination-activating gene 1 (RAG1), as per Gartner et al. (2013), the KIAA2018 ortholog (KIAA2018), following Portik et al. (2012), and the dynein axonemal heavy chain 3 (DNAH3), nerve growth factor beta polypeptide (NGFB) and synuclein alpha interacting protein (SNCAIP) following Townsend et al. (2008, 2011).

Sequences were edited with Geneious Pro 6 (Biomatters, Auckland), aligned with the Geneious algorithm, and deposited in Genbank (accession numbers KM204350-4, KM598666-749, and

KX760196-1163).

Heterozygous positions in nuclear genes were called with Geneious plugin Find

Heterozygotes, using a 0.90 overlap threshold. For coalescent-based analyses, the haplotypic phase of heterozygotes was determined using PHASE 2.1.1 (Stephens and Donnelly, 2003), with a 0.90 probability threshold and a parent-independent mutation model. Input files for PHASE were prepared in SeqPHASE (Flot, 2010). For phylogenetic analyses, models of nucleotide

50

evolution for each locus were determined with JModelTest 2.1 (Darriba et al., 2012) implementing the Bayesian information criterion for model selection (Sullivan and Joyce, 2005).

Phylogenetic inference

To characterize the historical relationships among samples, which may help to identify former routes of dispersal, we used the multi-locus dataset to generate phylogenetic trees under a coalescent framework, using the *BEAST tool in BEAST 1.8.3 (Drummond et al., 2012). To assign individuals to putative independently evolving lineages (sometimes referred to as

“species”), as required by *BEAST, we used a Bayesian implementation of the Generalized

Mixed Yule-Coalescent model (GMYC; Pons et al., 2006). GMYC models the splits on a phylogeny as either divergences between “species” or as coalescent events within “species”.

Based on the assumption that the rate of within-species coalescence is much larger than the rate of divergences between species, the method aims to find a threshold that distinguishes these two branching types on a target phylogeny. Its Bayesian implementation (bGMYC, Reid and

Carstens, 2012) accounts for phylogenetic uncertainty and error by integrating over the posterior distribution of trees from a Markov Chain Monte Carlo (MCMC) procedure. To delimit putative independently evolving lineages within our target lizard taxa, we implemented bGMYC on the highly structured mitochondrial gene trees, also inferred with BEAST. Specifically, bGMYC was applied to 500 posterior trees from the mitochondrial BEAST runs. Using the bgmyc.multiphyl() function in R 3.0.2 (R Core Team, 2017), we implemented 50,000 MCMC steps (with 40,000 steps as burn-in), sampling every 100 steps.

In all BEAST analyses (i.e., mitochondrial gene trees and multi-locus coalescent-based trees), we ran three independent chains of 100 million steps, sampling every 10,000 steps.

51

Analyses were performed for each target species separately and employed a strict molecular clock. Convergence and stationarity of model parameters were assessed in Tracer 1.6

(Drummond et al., 2012) to ensure effective sampling sizes > 200. Runs were combined in Log

Combiner 1.8.3, with 10% of each run discarded as burn-in. A maximum clade credibility tree was summarized with Tree Annotator 1.8 (Drummond et al., 2012). Resulting topologies were visualized in FigTree 1.4 (available from http://tree.bio.ed.ac.uk/software/figtree).

Testing alternative demographic models of dispersal across forest domains

We used coalescent simulations and Approximate Bayesian Computation (ABC) to test which Amazonian region (eastern vs. southwestern) most likely acted as the colonization source for the Atlantic Forest populations, and to test whether colonization was followed by a population bottleneck, as expected as a result of spatially or temporally restricted connections across domains. Our approach hence compared the observed (i.e., sampled) genetic data to data simulated through a coalescent framework under competing demographic scenarios. Both empirical and simulated sequences were recapitulated into informative summary statistics that capture patterns of genetic variation (Beaumont, 2010; Csilléry et al., 2010). For each of our three focal species, previous research found that Atlantic Forest samples compose a clade nested among Amazonian ones, suggesting a single invasion of the Atlantic Forest from Amazonia

(Prates et al., 2015; Prates et al., 2016a). We therefore tested alternative historical scenarios that differ by 1) which Amazonian lineages (eastern or southwestern) are more closely related to

Atlantic Forest samples, thus potentially indicating connection routes, and 2) whether there was a founder event during colonization of the Atlantic Forest (indicating a bottleneck followed by pronounced population expansion), or if population sizes remained relatively constant in each

52

forest block (indicating vicariant subdivision of a large ancestral population). Four scenarios were simulated and compared (Fig. 3.1). In scenario A (northeast vicariance), an ancestral population occurring in both eastern Amazonia and the northern Atlantic Forest was split following expansion of open and dry domains in northeastern South America. In scenario B

(northeast dispersal), a founding population colonized the northern Atlantic Forest from eastern

Amazonia through a forest corridor in the presently dry belt of northeastern South America, with subsequent population expansion within the Atlantic Forest. Scenarios C (southwest vicariance) and D (southwest dispersal) are similar to A and B, respectively, with the difference that southwestern (instead of eastern) Amazonia acted as the source of individuals into the Atlantic

Forest (Fig. 3.1). In the case of Polychrus marmoratus, we limited our analyses to scenarios A and B due to the lack of individual samples from southwestern Amazonia.

Individuals from the two better-sampled (Anolis) species were assigned to either an eastern or southwestern Amazonian group according to their location in relation to longitude

60oW (east or west). This longitude has been proposed as the approximate region where two major Amazonian macroclimatic systems meet (Cheng et al., 2013), and broadly corresponds to a known biogeographic break for several Amazonian taxa (e.g., d'Horta et al., 2013; Fernandes et al., 2013; Geurgas et al., 2010; Hal and Harvey, 2002). It also marks a region of contact for two genetic groups within the Amazonian Anolis punctatus, as inferred through a genetic clustering analysis (Prates et al., 2016a).

Previous clustering analyses of Anolis punctatus DNA sequences recovered a major break between eastern and western Amazonian samples, as well as two distinct genetic clusters in western Amazonia: one restricted around the Brazil-Peru border, and the other spanning a much larger range in central and southwestern Amazonia (Prates et al., 2016a). This pattern was

53

not observed in Amazonian A. ortonii, for which substructure was not detected (Prates et al.,

2016a). To ensure that the spatial groups used in our demographic analyses did not violate panmixia, and given the limited availability of samples for the A. punctatus cluster occurring at the Brazil-Peru border, our demographic analyses included only samples of the more broadly distributed western A. punctatus cluster (i.e., that occurs in central and southwestern Amazonia).

For each lizard species independently, we performed four million coalescent simulations

(one million simulations under each of the four demographic scenarios) followed by ABC inference with DIYABC 2.1 (Cornuet et al., 2014). For each locus, the following four summary statistics were computed: mean of pairwise genetic difference within groups, Tajima’s D

(Tajima, 1989), number of segregating sites between groups, and pairwise Fst (Weir and

Cockerham, 1984), totalizing 72 summary statistics (given the six sampled loci and three spatial groups). These summary statistics were chosen based on preliminary runs which confirmed that

1) they can discriminate among the tested scenarios and 2) the empirical data were contained within the space of simulated data as outlined by these summary statistics (Cornuet et al., 2014).

To verify that the observed genetic data were contained within the space of simulated data, we performed a principal component analysis (PCA) on the simulated summary statistics. Then, we assessed the statistical support of each scenario by calculating Euclidean distances between the observed and each simulated dataset, using a logistic regression on the summary statistics

(Beaumont et al., 2010). We estimated the posterior probability of each scenario based on the

2,000 simulated datasets (0.05% of the total simulations) that were closest to the observed data.

Posterior distributions of population parameters were estimated under the best-fit scenario, using the 1,000 simulated datasets that were closest to the observed data.

To evaluate the accuracy of the model selection procedure (i.e., to verify whether the

54

alternative scenarios tested can be discriminated based on summary statistics), we simulated

1,000 pseudo-observed datasets under each scenario, such that the true scenario was known for each pseudo-observed dataset. We used the summary statistics to calculate Euclidean distances between each pseudo-observed dataset and the simulated data, based on one million simulated datasets. Lastly, we calculated the proportion of pseudo-observed datasets correctly identified by the model selection procedure (Cornuet et al., 2014).

Priors for model parameters were set as follows: divergence times between regions ~ uniform [0.1, 5] million years, assuming a generation time of one year in anole lizards to convert prior values from number of generations to years (Jezkova et al., 2009; Munoz et al., 2013; Tollis et al., 2012); effective population size in each region ~ uniform [0.01, 5] million; effective population size during a founder event ~ uniform [0.001, 0.1] million; effective population size prior to a vicariant event ~ uniform [0.01, 5] million. An HKY85 substitution model (Hasegawa et al., 1985) was implemented for each gene. Mutation rate priors were set as ~ uniform [10-11,

10-9] per site per generation for the five nuclear loci, and as ~ uniform [10-9, 10-7] per site per generation for the mitochondrial locus. Prior ranges for mutation rates were based on the number of substitutions among sequences of Anolis and Polychrus species and on estimated divergence times between these species based on fossil calibrations (Prates et al., 2015). In each simulation, the time of Atlantic Forest colonization was constrained to be more recent then the time of coalescence between the eastern and southwestern Amazonian groups. Likewise, the time of a population bottleneck in the Atlantic Forest was constrained to be more recent than the time of coalescence between the Atlantic Forest group and its Amazonian source.

55

Testing synchronous dispersal between species across forest domains

After determining which Amazonian region was the most likely source of Atlantic Forest colonization, we used a hierarchical Approximate Bayesian Computation approach (hABC) to test whether colonization of the Atlantic Forest happened synchronously in these three species.

With that, we compared the sampled genetic data to data simulated under alternative divergence models differing in their combinations of synchronous or idiosyncratic divergence times between pairs of populations of distinct species. We used dpp-msbayes as implemented in the PyMsBayes package (Oaks, 2014), an extension of msbayes (Huang et al., 2011) that implements a Dirichlet process over the hyper-prior specifying the number of divergence events (Oaks et al., 2013). A total of two million simulations were performed, and the posterior probability of alternative divergence models was estimated based on the 1,000 (0.05%) simulated datasets that were closest to the observed data. Re-sorting of taxa and regression correction were not implemented in dpp-msbayes analyses (Oaks et al., 2013; Oaks, 2014).

Posterior estimates of demographic parameters from DIYABC analyses were used as a guide to set prior distributions in dpp-msbayes, yet modified to encompass broader intervals to better account for uncertainty and error in DIYABC estimates, as follows: tau ~ gamma [1, 0.0], based on estimates of mutation rates and divergence times between populations; theta ~ gamma

[1, 0.005], based on estimates of mutation rates and population sizes; and the magnitude of population bottlenecks ~ beta [1, 1], which allows a wide range of bottlenecks strengths, including no bottleneck. We allowed theta and the magnitude of bottlenecks to differ between each pair of populations. The substitution rate of the mitochondrial locus was set as 10 times faster than that of nuclear genes (from DIYABC estimates), and the transition-to-transversion rate ratio of the HKY substitution model (Hasegawa et al., 1985), implemented in dpp-msbayes,

56

was estimated for each locus of each species separately using JModelTest 2.1 (Darriba et al.,

2012). The concentration parameter of the Dirichlet process hyper-prior was set as ~ gamma

[1000, 0.0014], such that equal prior probability was assigned to divergence models where all or no taxa diverged synchronously. Higher prior probability was assigned to models with intermediate numbers of divergence events (as there are three possible models in which two out of three taxa co-diverge). These settings rendered our analysis conservative with regard to the occurrence of synchronous divergences between the three target species.

To test the hypothesis of divergence followed by gene flow between Amazonia and the

Atlantic Forest, a result expected if forest connections were recurrent through time, we compared the statistical fit of divergence models that incorporated migration to those that did not. In this case, the migration rate prior was set as ~ gamma [1, 0.01], while the remaining model parameter settings were kept the same.

To test whether there is a bias towards synchronous divergences in our dpp-msbayes analyses, we simulated 1,000 pseudo-observed datasets under a scenario of asynchronous divergences (i.e., enforcing three independent divergence times), which were then analyzed under the scenario used for inference (described above), based on 500,000 simulated datasets.

With that, we assessed the frequency of 1) the number of inferred divergence events (which was known to be three) and 2) the posterior probability of one synchronous divergence (which was known to be incorrect). Moreover, to assess the robustness of our estimates of the dispersal index of divergence times to the inclusion of a migration parameter, we simulated 1,000 pseudo- observed datasets under the scenario incorporating migration, which was then analyzed under the scenario of no migration based on 500,000 simulated datasets.

57

Results

Phylogenetic structure

The bGMYC analyses identified 10 putative independently evolving lineages within

Anolis punctatus, 17 in A. ortonii, and seven in Polychrus marmoratus. Most individual lineages were restricted to the Atlantic Forest, eastern Amazonia, or southwestern Amazonia. A single lineage encompasses all Atlantic Forest samples of A. punctatus (J in Fig. 3.2a); this lineage also includes samples from two sites in the eastern Amazonian border in the Brazilian states of Pará and Mato Grosso (PA and MT in Fig. 3.2, respectively). Atlantic Forest samples of A. ortonii composed four lineages, while three lineages represent P. marmoratus in this region.

For the three species, coalescent-based phylogenetic analyses recovered Atlantic Forest lineages as nested within Amazonian lineages, supporting the hypothesis that ancestral ranges trace back to Amazonia (Fig. 3.2). Moreover, in the case of Anolis punctatus and A. ortonii (the two species with available samples for all three regions), phylogenetic analyses found Atlantic

Forest lineages to be more closely related to those from eastern Amazonia than to those from southwestern Amazonia. Southwestern Amazonian lineages of A. punctatus compose a maximally supported clade (Posterior Probability (PP) = 1), while eastern Amazonian and

Atlantic Forest lineages compose another major clade (PP = 1). In the case of A. ortonii, the four

Atlantic Forest lineages compose a moderately supported clade along with one lineage from eastern Amazonia in the state of Pará (PP = 0.85). Similar to A. ortonii, Atlantic Forest

Polychrus marmoratus are paraphyletic (Fig. 3.2). However, for both A. ortonii and P. marmoratus, nodal support values for the relationships between lineages were generally low

(Fig. 3.2).

While phylogenetic relationships among Atlantic Forest lineages were poorly supported

58

in the coalescent-based analyses, mitochondrial gene trees further support the view that Atlantic

Forest samples are more closely related to the eastern than to the southwestern Amazonian samples (Appendix 3.2a). Furthermore, these trees nest the southern Atlantic Forest samples within those from the northern Atlantic Forest. In Anolis punctatus, for instance, a clade of individuals from northeastern Brazil, sampled in the state of Alagoas (AL in Fig. 3.2), is sister to all remaining Atlantic Forest samples (PP = 0.95). Within the latter group, a clade from the state of Bahia (BA in Fig. 3.2) is recovered as sister to a clade composed of samples distributed in more southern sites (PP = 1). Similar patterns were recovered within A. ortonii and Polychrus marmoratus (Appendix 3.2b,c).

Best-fit historical scenarios

Based on coalescent simulations and Approximate Bayesian Computation, we find that the observed genetic data are consistent with the idea that the better-sampled Anolis punctatus and A. ortonii colonized the Atlantic Forest from eastern Amazonia. For these lizards, alternative scenarios posing colonization of the Atlantic Forest from southwestern Amazonia received low to no support (PP < 0.03).

The best-fit demographic scenario differed between the three species. For Anolis punctatus (PP = 0.99) and A. ortonii (PP = 0.68), results favor the scenario of a population bottleneck associated with the colonization of the Atlantic Forest, followed by pronounced population expansion in the Atlantic Forest (Fig. 3.1b). On the other hand, the best-fit scenario for Polychrus marmoratus (PP = 0.81) was that of an ancestral population occurring in both eastern Amazonia and in the Atlantic Forest which was then split between domains, with no detectable population bottleneck involved (Fig. 3.1a).

59

Posterior parameter estimates based on best-fit scenarios for Anolis punctatus, A. ortonii, and Polychrus marmoratus support mid-Pleistocene divergences between Atlantic Forest and eastern Amazonian lineages, dating back to 545 kya (median value; 95% credibility interval

(C.I.) = 0.19-1.62 mya) in A. punctatus, 501 kya (C.I. = 0.16-1.24 mya) in A. ortonii, and 740 kya (C.I. = 0.14-2.37 mya) in P. marmoratus (posterior estimates of all model parameters in

Table 3.1; plots of parameter estimates presented in Appendix 3.3). Effective population size estimates suggest that the initial pool of Atlantic Forest colonizers was ca. 95-fold smaller than that of the ancestral eastern Amazonian population in A. punctatus and ca. 85-fold smaller in A. ortonii (Table 3.1), supporting a pronounced population bottleneck following dispersal into the

Atlantic Forest in the two Anolis. The DIYABC results also suggest that A. punctatus and A. ortonii underwent an 89-fold and 35-fold population expansion in the Atlantic Forest, respectively, once it was colonized. By contrast, no population bottleneck or subsequent population expansion was detected in the Atlantic Forest samples of P. marmoratus (Table 3.1).

Median substitution rates were estimated as 1.27 x 10-8, 1.72 x 10-8 and 8.2 x 10-9 substitutions per site/year for the mitochondrial locus of A. punctatus, A. ortonii, and P. marmoratus, respectively, and ranged between 1.31 x 10-10 and 7.92 x 10-10 substitutions per site/year for each nuclear marker of the three species (see Appendix 3.4 for substitution rates estimated for all loci).

Model validation based on principal component analyses of the summary statistics confirmed that the observed data were contained within the space of simulated data, in all three species (Appendix 3.3). Posterior error rates based on pseudo-observed datasets were estimated as 0.19 for both Anolis punctatus and A. ortonii, and 0.09 for Polychrus marmoratus, suggesting that simulated scenarios were overall mutually identifiable.

60

Best-fit co-divergence models

We found support for synchronous divergences between eastern Amazonian and Atlantic

Forest lineages across Anolis punctatus, A. ortonii, and Polychrus marmoratus (PP = 0.48). We recovered significantly lower support for a scenario of completely temporally-idiosyncratic divergences (PP < 0.09), as well as for each of the three scenarios in which two out of the three species have co-diverged (PP < 0.17). This result is further supported by the dispersion index of divergence times (the variance/mean of the divergence times across all population pairs), which equaled zero (C.I. = 0-0.83). In agreement with the DIYABC analyses, posterior estimates from dpp-msbayes suggest that synchronous divergences between Amazonian and Atlantic Forest populations of the three species happened at around 950 kya (C.I. = 0.20-2.08 mya), assuming an average substitution rate across loci of 2.42 x 10-9 per generation (from DIYABC estimates).

An analysis incorporating post-divergence migration between eastern Amazonian and

Atlantic Forest populations received slightly lower support than one that did not incorporate gene flow, yet posterior probabilities were very similar between models (PP = 0.48 vs. 0.52). This result suggests that the genetic data (or the summary statistics implemented in dpp-mrbayes) do not provide enough signal to capture migration across forest blocks following initial population divergence.

Validation analyses based on pseudo-observed datasets confirm that our dpp-msbayes analyses were not biased towards a scenario of synchronous divergences (Appendix 3.5).

Validation analyses also confirm that our estimates of the dispersal index of divergence times are robust to the incorporation of a migration parameter (Appendix 3.5).

61

Discussion

This study provides insights about likely routes of former forest connections and the directionality of dispersal between Amazonia and the Atlantic Forest. An Amazonian origin for

Anolis ortonii, A. punctatus, and Polychrus marmoratus agrees with patterns of phylogeographic structure previously recovered for birds (Batalha-Filho et al., 2013), rodents and marsupials

(Costa, 2003), and other lizards (Pellegrino et al., 2011). Our findings support the hypothesis that past forest expansion in northeastern South America promoted forest connections and enabled biotic exchange between these two major forest blocks (Cheng et al., 2013). They also reinforce the view that climatic variation over time has led to taxonomically similar and historically linked communities between eastern Amazonia and the northern component of the Atlantic Forest.

Our investigation sheds light on the timing of historical biotic exchange between

Amazonia and the Atlantic Forest, and supports the hypothesis that documented precipitation shifts during the Quaternary have been key drivers of forest cover change in South America

(Cheng et al., 2013). The conclusion that the three focal lizard species expanded synchronously into the Atlantic Forest in the mid-Pleistocene is consistent with paleoenvironmental data available; patterns of travertine deposition suggest pulses of rainforest expansion over the xeric

Caatinga as much as 900 kya (Auler et al., 2004), which is close to the range of colonization times that we estimated for Anolis punctatus, A. ortonii, and Polychrus marmoratus. Because the amplitude of recurring precipitation increase in northern South America varied over time, matching the precession component of Earth’s orbital cycles (Cheng et al., 2013), it is possible that connections and biotic exchange between Amazonia and the Atlantic Forest were restricted to (or more pronounced at) time periods in which levels of precipitation were particularly high.

Our data suggest that one such pivotal event may have occurred around 950 thousand years ago.

62

A scenario of recurrent forest connections could have potentially led to periodic re- establishment of gene flow after initial population separation. Paleontological and geochemical studies have inferred pulses of rainforest expansion around present-day northeastern Brazil happening as recently as the Holocene (Cheng et al., 2013; de Vivo and Carmignotto, 2005; de

Oliveira et al., 1999). However, our tests of gene flow following initial divergence were inconclusive, providing equal support for historical models that incorporated post-divergence migration and models that did not. Intriguingly, phylogenetic patterns in a range of taxa indicate no repeated dispersal events between Amazonia and the Atlantic Forest following initial colonization (e.g., Fouquet et al., 2012a,b; Geurgas and Rodrigues, 2010; Pellegrino et al., 2011;

Rodrigues et al., 2014), in spite of the opportunity provided by subsequent periods of increased humidity and presumed forest expansions.

Due to the cyclical nature of Late Quaternary climate change (Cheng et al., 2013), one could expect to detect signals of population size shifts in the Atlantic Forest after colonization.

However, our analyses estimated that the pulse of population expansion in Atlantic Forest Anolis punctatus and A. ortonii (Table 3.1) precedes multiple subsequent cycles of climate change documented for this region (Cheng et al., 2013). The fact that we detected no signatures of more recent population shifts suggests that patterns of genetic diversity of the two Anolis were strongly shaped by the initial colonization event. However, more recent demographic changes have been inferred for these three species in the Atlantic Forest based on population-level analyses using reduced genomic data (Prates et al., 2016a). That study recovered population expansion in A. punctatus around 60 kya, which is later than the time of expansion estimated from our DIYABC analyses of this species (~290 kya). In the case of A. ortonii and Polychrus marmoratus, Prates et al. (2016a) found signatures of population contraction in the Atlantic Forest around 72 and 81

63

kya, respectively, which contrasts with population expansion in A. ortonii (~276 kya) and no size change in P. marmoratus based on our DIYABC analyses. Given the different estimated times and population trends (in the case of A. ortonii and P. marmoratus) between Prates et al. (2016a) and the present investigation, it seems likely that these two studies captured distinct and consecutive demographic events. If so, it is possible that the genetic datasets used (i.e., six coding regions vs. thousands of single nucleotide polymorphisms) provide signals of distinct demographic events which have occurred at different time scales. Nevertheless, it is currently unclear whether and how results based on traditional and next-generation genetic datasets are comparable.

We find that anole lizards do not fit to a scenario of colonization through former forest corridors around present-day southwestern Brazil, which have been implicated in instances of closely related taxa occurring in both southern Atlantic Forest and southwestern Amazonian or

Andean forests (Batalha-Filho et al., 2013; Cheng et al., 2013). Based on divergence time estimates among bird taxa, it has been suggested that southern connections between these two rainforest blocks generally predate 5.6 mya, whereas northern connections were more recent, between 5.5 and 4.17 mya (Batalha-Filho et al., 2013). Old divergences (dating to the Miocene and Oligocene) have been also reported for a number of lizard clades that show disjunct distribution across forests, such as between species of Enyalius (Rodrigues et al., 2014) and between the sister genera Leposoma and Loxopholis (Pellegrino et al., 2011) and Chatogekko and

Coleodactylus (Geurgas and Rodrigues, 2010). The same was found between frog species of

Adelophryne (Fouquet et al., 2012) and Adenomera (Fouquet et al., 2014), and between the genera Dendrophryniscus and Amazophrynella (Fouquet et al., 2012). However, spatial patterns of phylogenetic structure suggest that most of these old divergences represent northern

64

colonizations, similar to our recently diverged populations. This pattern disagrees with the idea that different dispersal routes were restricted to distinct time scales (Batalha-Filho et al., 2013).

Lastly, our results indicate that the expansion of suitable habitats and associated opportunities for dispersal resulted in distinct demographic patterns among species. While we recovered remarkable temporal congruence in the colonization of the Atlantic Forest between the focal lizards, the best-fit historical demographic scenarios differed among them. ABC analyses on the observed genetic data are consistent with a history of pronounced population expansion in

Anolis punctatus and A. ortonii after colonization of the Atlantic Forest. By contrast, results for

Polychrus marmoratus are consistent with the subdivision of a large, panmictic ancestral population. The effective population size of P. marmoratus in the Atlantic Forest at the time of colonization was estimated as two orders of magnitude larger than that of the two other species in the same region (Table 3.1). These findings agree with the view that former reconfiguration of

South American rainforests had species-specific impacts on the local fauna (Prates et al., 2016a).

We hypothesize that the contrasting patterns observed across these three species may be associated with distinct tolerances to forest fragmentation and to a matrix of open and drier environments (Caatinga and Cerrado). This idea is consistent with available information on the ecology and distribution of these species, as P. marmoratus has been recorded in transition zones between Amazonian forests and Cerrado savannas, while A. ortonii and A. punctatus seem to be strictly associated to closed-canopy wet forests (Kawashita-Ribeiro and Ávila, 2008; Ribeiro-

Júnior, 2015; Vitt et al., 2003). In the case of P. marmoratus, some capacity to tolerate transition areas between South American rainforests and their adjacent settings may have enabled higher levels of gene flow under environmental conditions that were rather restrictive for the two Anolis species. This hypothesis has profound implications for biogeographic studies: despite the

65

magnitude of corridors between regions, the apparent biogeographic history of a species may be strongly determined by its dispersal capacity within and across patches of suitable habitat, which could confound the distinction between dispersal and vicariance.

Based on ample geographic sampling of genetic diversity in three co-distributed lizard species, we track historical relationships between populations and the timing of former biotic exchange between highly diverse tropical rainforest domains. The spatial patterns of genetic structure indicate that cessation of environmental breaks as a result of inferred climatic fluctuations across time provided key opportunities for dispersal and species exchange between regions. Intriguingly, we found this influx of species to be asymmetric, in agreement with the hypothesis that some regions have acted mostly as providers and others as receivers of biodiversity (Santos et al., 2009). Our results also point to species-specific responses to the former expansion of suitable habitats, strengthening assertions that biological attributes may play an important role by shaping responses to shared environmental change (Paz et al., 2015;

Zamudio et al., 2016). Understanding whether and how climate change over time has led to opportunities for dispersal and biotic exchange, and how different taxa have explored such opportunities, will contribute to investigations of community assembly in other biologically diverse systems.

66

CHAPTER 4: Genomic divergence and local adaptation during range expansions in

widespread South American anoles

Introduction

Environmental variation across geographic space limits species ranges, therefore affecting ecological processes and large-scale biogeographic patterns (Navas, 2002). Abiotic factors such as temperature and water availability, for instance, strongly constrain physiological function and associated behavior, especially in ectotherms (Barros et al., 2010; Prates et al.,

2013). Widely distributed species, which frequently span heterogeneous landscapes, must deal with these constraints. In widespread organisms, natural selection may favor alleles that increase survival and reproduction under certain environmental conditions, leading to local adaptation

(Kawecki and Ebert, 2004). However, the potential for local adaptation can be limited by the homogenizing effects of gene flow between populations (Bridle and Vines, 2007; Lenormand,

2002; Sexton et al., 2014). Uncovering whether environmental gradients are linked to the genetic makeup of organisms, while documenting how gene flow constrains population differentiation, can greatly improve our understanding of the process of local adaptation and its implications to genetic divergence across ecological gradients (Nosil, 2012; Rellstab et al., 2015).

Because genetically determined physiological tolerances restrict species dispersal

(Ghalambor et al., 2006), local adaptation may play a major role in the establishment of organisms in new environments (Hohenlohe et al., 2010; Huey et al., 2000). It has been suggested, for instance, that range expansions require adaptation to conditions at or just beyond the range edge (Bridle and Vines, 2007). However, experimental investigations support the view that physiological tolerances are generally highly conserved within species over wide

67

environmental gradients, as in many ectothermic vertebrates (e.g., Crowley, 1985; Hertz et al.,

1983; John-Alder et al., 1989; Prates et al., 2013; Van Damme et al., 1989). As a result, it is unclear the extent to which expansions into new environments rely on physiological adaptation and underlying genetic differentiation, as opposed to, for instance, exaptation (“pre-adaptation”;

Gould and Vrba, 1982) or phenotypic plasticity (which, although potentially adaptive, does not imply genetic divergence between populations; Buskirk, 2009). Studies of genetic differentiation among closely related lineages that occupy contrasting habitats can therefore shed light on the role of local adaptation for the establishment in new habitats and geographic regions (Campbell-

Staton et al., 2016).

In this investigation, we examine potentially adaptive intra-specific genetic variation in the context of range expansions over wide environmental gradients. For that, we focus on populations of two South American anole lizards, Anolis ortonii and Anolis punctatus. The distributions of these two distantly related co-distributed species, which diverged around 49 million years ago (Prates et al., 2015, 2017), span a wide ecological gradient over more than 25 degrees of latitude over most of Amazonia and the coastal Brazilian Atlantic Forest (Ribeiro-

Júnior, 2015). Previous studies have suggested that they have undergone range expansions, such as colonization of the northern Atlantic Forest from eastern Amazonia, and posterior dispersal into the southern Atlantic Forest (Prates et al., 2016a,b). Our comparative study has two main goals. By sampling reduced genomic data of these species, we first seek to track the history of range expansions in each anole. Through historical demographic inference under a coalescent framework, we assess gene flow among populations from distinct habitats within each species’ range - a force that could oppose local adaptation. To evaluate whether species occurrence in diverse environments is associated to local adaptation, we then implement genome-environment

68

association analyses. We examine associations between allele frequencies in thousands of loci and spatial gradients of temperature and precipitation across each species’ range. For that, we control for the effects of demographic history, as inferred through patterns of genetic structure, because it can lead to allele frequency patterns that mimic signatures of selection (Rellstab et al.,

2015). By performing functional annotations in the resulting candidate loci (i.e., loci found to be linked to environmental variation across sampled localities), we identify genes that may have favored the establishment of each species in diverse habitats, as well as their associated biological processes.

South American anole lizards are well suited for this investigation. Natural history information indicate that these lizards do not rely on behavioral thermoregulation (Vitt et al.,

2002, 2003), which points to a role of genetically-determined physiological traits in shaping ecological tolerances (e.g., Navas, 2002). Accordingly, studies on Caribbean anoles suggest that physiological performances and thermal niches evolve rapidly in this group of lizards (Kolbe et al., 2012; Leal and Gunderson, 2012; Muñoz et al., 2014). Importantly, while Caribbean anoles are central to studies of adaptive radiation and evolutionary ecology, we know very little about the biology and evolution of mainland species, which effectively correspond to most of the taxonomic and ecological diversity of the anole family (Losos, 2009).

Material and Methods

Molecular sampling

We generated reduced genomic data for 47 individuals of Anolis punctatus and 23 of A. ortonii, building upon the sampling of Prates et al. (2016a) (voucher and locality information presented in Appendix 4.1). Genomic DNA was extracted from liver or tail fragments preserved

69

in 100% ethanol through a high-salt extraction protocol following proteinase and RNAase treatment. After ensuring that the DNA was not degraded using agarose gels, DNA concentration was measured in a Qubit 2 fluorometer (Invitrogen, Waltham) to ensure a final concentration of

30-100 ng DNA/ul in a total volume of 30 ul (in TE buffer). A restriction-site associated DNA library was generated through a genotype-by-sequencing protocol (Elshire et al., 2011) at the

Institute of Biotechnology at Cornell University. Briefly, genomic DNA was reduced to a representation library using the EcoT22I restriction enzyme, and the resulting fragments were tagged with individual barcodes, PCR-amplified, multiplexed, and sequenced in an Illumina

HiSeq platform. The number of single-end reads per individual ranged from around 500,000 to six million.

We used the Ipyrad 0.6.17 pipeline (available at http://ipyrad.readthedocs.io) to de- multiplex and assign reads to individuals based on sequence barcodes (allowing no nucleotide mismatches from the barcode of each individual), to perform de novo read assembly (clustering similarity threshold > 0.9), to align reads into loci, and to extract single nucleotide polymorphisms (SNPs). To filter out poor-quality reads and reduce base-calling error, a minimum Phred quality score (= 33) and minimum sequence coverage (= 10x) were enforced. To reduce potential paralogs, we imposed a maximum proportion of heterozygous sites per locus (=

0.25), a maximum number of heterozygous individuals per locus (= 5), and a maximum number of SNPs per locus (= 10), while ensuring that variable sites were biallelic. Final locus length ranged from ~50-140 base pairs. For phylogenetic and genetic clustering analyses, a single SNP was randomly extracted per locus to reduce linkage disequilibrium across sites and to maximize sampling of independent SNP histories (thereafter referred to as unlinked SNPs). Genetic datasets used in all analyses were deposited in Dryad (pending article submission).

70

Inferring population genetic structure

To approximate the number of populations in Anolis ortonii and A. punctatus, we estimated genetic clusters within each species separately using sNMF (Frichot et al., 2014). sNMF estimates ancestry coefficients for each sample based on an entropy criterion, which evaluates the statistical fit of the clustering model to the genetic data through a cross-validation procedure (“masking”; Frichot et al., 2014). Several values of the number of genetic clusters (k) were tested, with k ranging from 1 to 10. For each k, 50 replicates were run. The run with the lowest entropy value, estimated by masking 5% of the samples, was considered to identify the best k (Frichot et al., 2014). To examine the robustness of sNMF results given the choice of regularization parameter, we implemented regularization values = 1, 10, 100, and 1,000. Up to

30% of missing data were allowed in sNMF analyses (i.e., each DNA site was present in at least

70% of sampled individuals). The resulting dataset was composed of 18,849 unlinked SNPs for

A. ortonii and 11,463 for A. punctatus.

Inferring population trees

Using the unlinked SNP data, we inferred phylogenetic relationships between the genetic clusters identified by the sNMF analysis. For that, we used the coalescent-based method SNAPP

(Bryant et al., 2012) as implemented in BEAST 2.4.5 (Bouckaert et al., 2014). SNAPP analyses were performed for Anolis punctatus only, for which several genetic clusters were recovered; in the case of A. ortonii, where only two genetic clusters were found (see Results), we did not implement SNAPP and treated the two genetic clusters as sister to each other.

For the SNAPP analyses, a broad gamma-distributed prior, given by parameters shape (α) and rate (β), was implemented for the theta parameter (= 4Nμ, where N corresponds to the

71

effective population size and μ to the per-generation mutation rate), using α = 2 and β = 10.

Moreover, uniformly distributed priors were implemented for the mutation parameters, with an interval = 0-100. A uniform prior distribution with an interval = 400-700 was set to the rate of the Yule diversification process (lambda), given an expected tree height of ~0.002 to 0.003 substitutions per site based on a recent phylogeographic study that included Anolis punctatus

(Prates et al., 2016b). Two independent SNAPP runs of one million generations each were performed, with a sampling frequency of 1,000 steps and discarding 10% of generations as burn- in. After ensuring convergence between runs and stationarity of model parameters (effective sample sizes > 200), using Tracer 1.6 (Drummond et al., 2012), posterior tree densities were visualized in DensiTree 2.4.5 (Bouckaert et al., 2014). Due to computational constraints, population trees were inferred with a subset of the unlinked SNPs of A. punctatus (= 3,000) and a maximum of five individuals from each of the genetic clusters inferred by sNMF, allowing a maximum of 10% of missing data.

Estimating historical demographic parameters

To estimate population parameters such as coalescent times between populations and effective population sizes (current and ancestral), we used the full-likelihood method G-PhoCS

(Gronau et al., 2011). G-PhoCS estimates historical demographic parameters under a coalescent framework based on the entire sequence of loci (as opposed to SNPs only). For G-PhoCS analyses, populations of Anolis ortonii and A. punctatus were defined as per the genetic clustering inferred through sNMF. The historical relationships between clusters (i.e., the tree topology) of A. punctatus were set based on SNAPP results, while the two clusters found within

A. ortonii were treated as sister populations (Fig. 4.1). Because we are interested in assessing the

72

presence of historical gene flow between populations that occur in different habitats, we included, in each species’ model, migration bands between the Atlantic Forest cluster and its most closely related Amazonian cluster (with one migration band in each direction; Fig. 4.1). We performed G-PhoCS analyses for each anole species separately, implementing a Markov Chain

Monte Carlo of 100,000 generations and sampling every 100 generation. Due to computational constraints, only 3,000 loci of each anole species were used, allowing for a maximum of 10% of missing data. We assessed the stationarity of each run, as well the posterior distribution of model parameters, using Tracer 1.6 (Drummond et al., 2012). To estimate the posterior distribution of model parameters, 10% of each run was discarded as burn-in.

In G-PhoCS analyses, gamma-distributed priors, given by parameters shape (α) and rate

(β), were applied for the divergence times between populations and for effective population sizes, setting α = 1 and β = 1000 (for divergence time and population size priors). Additionally, the prior distribution of each migration band (given by the proportion of individuals in one population that arrived from another population in each generation) was set as a gamma distribution with α = 0.0002 and β = 0.00001. To convert parameter values from number of generations to years, we assumed a generation time of one year in anole lizards, following

Jezkova et al., 2009; Muñoz et al., 2013; Tollis et al., 2012. To convert mutation rate-scaled parameter estimates to absolute effective population sizes and divergence times, we assumed an average substitution rate of 2.4 x 10-9 substitutions per site per year, as estimated from a multi- locus study of mainland anoles (Prates et al., 2016b).

Choice of environmental variables

To describe environmental factors in a genome-environment association analysis, we

73

used bioclimatic variables from the WorldClim database (Hijmans et al., 2005) to summarize patterns of temperature and precipitation variation. WorldClim values were extracted from the collection site of each lizard sample. We limited our analysis to environmental variables that are likely ecologically relevant for anoles (reviewed by Losos, 2009), easily interpretable (i.e., no composite bioclimatic variables were used), and which showed clear geographic variation across the distribution of our study species, as follows: annual mean temperature (BIO1), maximum temperature of the warmest month (BIO5), mean temperature of the warmest quarter (BIO10), annual precipitation (BIO12), precipitation of the wettest month (BIO13), and precipitation of the wettest quarter (BIO16). Because temperature and precipitation variables are often highly correlated (i.e., pairwise Pearson correlation coefficients >0.7), we implemented principal component analyses (PCA) for each anole species separately based on all six environmental variables, and then used the first principal component in genome-environment analyses. This component explains >71% of the total environmental variation across sampled sites for each anole species. PCA was implemented using the prcomp function in R 3.4.0 (R Core Team,

2017).

Performing genome-environment association analyses

To test for associations between environmental descriptors and allele frequencies, we used latent factor mixed linear models (LFMM, Frichot et al., 2013). LFMM is a hierarchical

Bayesian model in which the effect of environmental variables and the effect of neutral genetic structure on allele frequencies are estimated jointly; neutral structure is modeled in the form of unobserved (latent) factors, the number of which can be informed based on the results of a genetic clustering analysis. LFMM has low rates of false positives and is robust to isolation-by-

74

distance and to a variety of sampling designs and underlying demographic histories (Rellstab et al., 2015).

To define the number of latent factors to be used in the LFMM analyses, we explored a range of values adjacent to the best number of genetic clusters (k) inferred by sNMF, as recommended by Frichot et al., 2013. Specifically, we performed preliminary runs incorporating one to six latent factors, and assessed whether the confounding effects of neutral genetic structure were properly controlled through inspection of histograms of corrected p-values

(Frichot et al., 2015). Three and five latent factors provided better control of the effects of neutral genetic structure of A. ortonii and A. punctatus respectively, and were therefore used in final LFMM analyses.

LFMM was implemented using the lfmm function of the package LEA (Frichot et al.,

2015) in R 3.4.0 (R Core Team, 2017), with 10 replicates of 10,000 iterations each and 5,000 iterations as burn-in. A maximum false-discovery rate of 0.01 was used. To maximize the number of SNPs included in these analyses, we excluded three Anolis punctatus samples that had less than 2,500 genotyped loci in the final dataset (LSUMZH 12577, MPEG 26102, and MTR

21474); all samples of A. ortonii were used. Previous to LFMM analyses, we used vcftools

(Danecek et al., 2011) to filter out non-biallelic SNPs, as well as SNPs whose minor allele frequency was lower than 0.05 (as recommended in the LFMM manual). A maximum of 30% of missing data was allowed in LFMM analyses (i.e., each SNP was present in at least 70% of sampled individuals. However, to reduce the potential effects of missing sites, we performed haplotype imputation based on the ancestry coefficients estimated by sNMF, using the impute function of the LEA package (as recommended by LFMM developers). After these filtering steps, the genetic dataset was composed of 23,641 SNPs of A. ortonii in 12,449 loci, and 20,639

75

SNPs of A. punctatus in 9,824 loci.

Annotating gene ontologies

LFMM analyses recovered several SNPs whose frequencies are linked to each environmental variable (see Results). To examine whether the loci harboring these SNPs correspond to genes, we ran BLAST+ (Camacho et al., 2008) to compare each sequence with the protein-coding regions of the genome of the green anole, Anolis carolinensis (Alföldi et al.,

2011), using the blastx algorithm. For the blasting step, a local database composed of protein sequences of A. carolinensis was extracted from NCBI's protein database (nr) based on a list of identification numbers (e.g., NCBI's GI numbers) of sequences corresponding to genes in this species. For blasting, we used a maximum e-value (which describes the number of hits expected by chance, with lower values indicating higher confidence in sequence matching) of 0.001. To examine the identity and biological function of genomic regions linked to environmental variation, we imported blast results into Blast2Go (Conesa et al., 2005) and performed gene mapping followed by annotation of gene ontologies.

Results

Population genetic structure and demographic history

Clustering (sNMF) analyses recognize the Atlantic Forest samples as a distinct genetic group from Amazonian samples in both Anolis ortonii and A. punctatus, suggesting genetic differentiation across forest blocks (Fig. 4.2). Within Amazonia, however, the patterns of genetic structure differ between species: while four genetic clusters were inferred for A. punctatus in

Amazonia, only one cluster was recovered for A. ortonii in this region. In the case of A.

76

punctatus, one such genetic group is widely-distributed in central and eastern Amazonia (cluster eam in Fig. 4.2), one is restricted to western Amazonia, near the Brazil-Peru border (wam1 in

Fig. 4.2), and two clusters occur by the Brazil-Bolivia border, roughly on each bank of the

Madeira River (wam2 and wam3 in Fig. 4.2). In the case of A. ortonii, the single genetic cluster found in Amazonia shows a broad distribution throughout this forest domain.

Phylogenetic relationships among genetic clusters of A. punctatus, inferred using SNAPP, nest the Atlantic Forest cluster within Amazonian samples (Fig. 4.2). Specifically, the Atlantic cluster is more closely related to the Amazonian cluster whose distribution spans central and eastern Amazonia (cluster eam in Fig. 4.2). The ancestor of these two sister groups is sister to the genetic cluster restricted to western Amazonia on the Brazil-Peru border (wam1). The two A. punctatus clusters that are restricted to the region are sister to each other (wam2 and wam3; Fig. 4.2).

G-PhoCS analyses suggest large effective population sizes, spanning hundreds of thousands to millions of individuals in all regions and both species, for both current and ancestral populations (Table 4.1). Median coalescent times between the Atlantic Forest cluster and its phylogenetically closest Amazonian cluster were estimated as around 313 kya for Anolis ortonii

(95% highest posterior density [HPD] = 266-354 kya), and as around 233 kya for A. punctatus

(HPD = 200-271 kya). In both lizards, levels of gene flow between clusters in distinct regions were estimated by G-PhoCS to be very low, approaching zero (Table 4.1). For instance, in A. ortonii, the median proportion of individuals in the Atlantic Forest that originated from its

Amazonian sister cluster was estimated as 1.03 x 10-11 per generation, while the proportion of individuals in Amazonia that originated from the Atlantic Forest cluster is around 5.35 x 10-12 per generation. Similarly, for A. punctatus, the median proportion of individuals in the Atlantic

77

Forest that originated from its most closely related Amazonian cluster was estimated as 1.49 x

10-11 per generation, while the proportion of migrants in the opposite direction was found as 1.08 x 10-11 per generation.

Genome-environment analyses

Genome-environment association analyses of Anolis ortonii, using LFMM, recovered 526 loci whose frequency is significantly correlated (at the 0.01 level) to the first principal component extracted from temperature and precipitation variables at collection sites. From those candidate loci, only 12 successfully blasted against protein-coding regions of the genome of A. carolinensis, and eight were successfully annotated for gene ontologies (Table 4.2). In the case of A. punctatus, allele frequencies at 639 loci are significantly linked to environmental variation, of which only 21 blasted against protein-coding regions in the A. carolinensis reference genome, with 14 being successfully annotated (Table 4.2).

Based on gene ontology annotations, the resulting candidate loci are generally involved with similar biological activities in both A. ortonii and A. punctatus. At a molecular level, these genes are associated with substrate-specific transporter activity, hydrolase activity, DNA-binding transcription factor activity, and biological molecule binding activity (protein binding, ion binding, organic cyclic compound binding, heterocyclic compound binding) in both species (Fig.

4.3). Similarly, at a higher functional level such as biological process (i.e, activities that affect entire cells or organisms), candidate loci in both A. ortonii and A. punctatus are involved with metabolic processes, regulation of biological processes, and response to stimuli, among others

(Fig. 4.3). Despite these general similarities, only one candidate locus, the predicted gene of

Mucin-2, was found to be significantly associated to environmental variation in both A. ortonii

78

and A. punctatus.

Discussion

Based on reduced genomic data of two widespread mainland anole species, we find support for a scenario of biotic interchange between Amazonia and the Atlantic Forest through forest corridors in the Pleistocene. We also find evidence of associations between pronounced environmental gradients in South America and the patterns of allele frequency shown by eight known protein-coding loci in A. ortonii and 14 in A. punctatus. These results are consistent with a scenario of local adaptation and associated genetic differentiation associated with range expansions across heterogeneous landscapes in these widely distributed taxa. Population genetic divergence due to natural selection may have been favored by the large effective population sizes of these anoles, as well as low migration rates across regions, as estimated by the historical demographic analyses.

Our divergence time estimates between Amazonian and Atlantic Forest populations of A. ortonii and A. punctatus indicate Pleistocene connections between the two forests, in agreement with the available geological evidence and previous phylogeographic investigations.

Paleontological and geochemical records support a history of pulses of increased precipitation and wet forest expansion in South America during the Quaternary, which have been implicated in intermittent connections between Amazonia and the Atlantic Forest (Cheng et al., 2013; de

Oliveira et al., 1999; de Vivo and Carmignotto, 2004). A previous multi-locus study of Anolis and Polychrus lizards also supported the existence of a forest corridor in present-day northeastern Brazil that favored synchronous, multi-taxa colonization of the Atlantic Forest from eastern Amazonia during the mid-Pleistocene (Prates et al., 2016b).

79

Our analyses suggest that post-divergence gene flow between populations from

Amazonia and the Atlantic Forest has been negligible for both Anolis ortonii and A. punctatus, in disagreement with an expectation of multiple pulses of gene flow based on the paleoenvironmental data. Recurrent forest connections, which have been proposed based on paleontological and geological evidence (e.g., Cheng et al., 2013; de Oliveira et al., 1999; de

Vivo and Carmignotto, 2004), could have led to periodic re-establishment of population gene flow across major forest blocks. An investigation aiming to test this idea based on a multi-locus dataset from lizard species found inconclusive results, likely due to limited signal in the genetic data available then (Prates et al. 2016b). Here, by implementing historical demographic inference based on thousands of genetic markers, we found evidence of limited exchange of migrants following initial colonization of the Atlantic Forest by anoles, despite the presumed opportunities provided by cyclic periods of increased humidity and associated forest expansions (Cheng et al.,

2013).

Low levels of gene flow between populations, and large effective population sizes in

Anolis ortonii and A. punctatus, may have favored local adaptation across ecological gradients.

In our historical demographic analyses, estimates of migration rates between populations approached zero, while effective population sizes were found to be around hundreds of thousands to millions of individuals. Several studies have shown that the potential for local adaptation in natural populations can be greatly limited by the homogenizing effects of gene flow

(Bridle and Vines, 2007; Lenormand, 2002; Sexton et al., 2014). Moreover, it is well established that small population sizes strongly constrain the capacity for adaptation in response to natural selection (Bridle and Vines, 2007). Low migration rates and large effective populations in A. ortonii and A. punctatus may have contributed to the establishment of allele frequency

80

differences across regions, some of which may have arisen in response to selection related to local environmental regimes.

Genome-environment association analyses support genetic differentiation at several protein-coding loci across pronounced environmental gradients in both Anolis ortonii and A. punctatus, which is consistent with a scenario of local adaptation. Studies of population pairs in contrasting habitats have found differences in the frequency of alleles that underlie ecologically- relevant processes (e.g., Akashi et al., 2016; Dennenmoser et al., 2016; Guo et al., 2016), yet fewer studies have tested whether such differences correlate to environmental predictors. The candidate genes recovered by our analyses are involved with the regulation of a number of biological processes, including gene transcription (Scm-like protein with four MBT domains 1,

Tyrosine-phosphatase non-receptor protein type 21), cell proliferation and differentiation

(Abnormal spindle-like microcephaly-associated protein, Stimulated by retinoic acid gene 8 homolog, Tyrosine-phosphatase non-receptor protein type 21), and organ development and growth (Abnormal spindle-like microcephaly-associated protein, Scm-like protein with four MBT domains 1, Stimulated by retinoic acid gene 8 homolog). Studies in anoles have shown that growth and development are strongly influenced by abiotic factors such as temperature, with impacts on organismal fitness (Losos, 2009). In the case of A. ortonii and A. punctatus, selection for genetic variants that help sustain cell and organismal function under local environmental regimes may have been instrumental for range expansions. A similar scenario has been suggested for the North American green anole, A. carolinensis, in which colonization of colder regions has been linked to adaptive shifts in genes related to development, energy metabolism, and immune response (Campbell-Staton et al., 2016).

In both Anolis ortonii and A. punctatus, a significant association is observed between

81

environmental predictors and allele frequency variation in Mucin-2. The protein encoded by this gene is involved with the protection of the intestinal epithelium by forming an insoluble mucous barrier; it has a demonstrated role in the prevention of bacterial infections, while its down- regulation has been implicated in pathologies of the digestive tract (Xue et al., 2014). This result may suggest that environmental gradients may have favored alleles that affect the digestive function of anoles, in agreement with physiological studies that support strong effects of temperature on lizard digestion (Harwood, 1979). On the other hand, this association may reflect other factors that co-vary with temperature and precipitation across space, such as the composition of parasite or arthropod prey assemblages (e.g., Fallon et al., 2005).

Although we find a substantial number of loci linked to environmental variation in both anole species, only a few match those in the reference database of A. carolinensis protein-coding regions (Alföldi et al., 2011). Blasting could have been impaired by the short length of the restriction enzyme-associated loci that were sequenced in this study. Alternatively, several of the loci that we found to be associated to environmental variation may not effectively correspond to protein-coding genes, but instead to regions involved in gene expression (e.g., enhancers, transcription factor recognition sites) or that are physically linked to the genes under selection.

We argue, however, that the evolutionary distance between A. carolinensis, whose genome served as a reference, and the two anoles studied here has likely limited our inferences. The most recent common ancestor (MRCA) of A. carolinensis and A. ortonii has been estimated as around

38 million years old, while the MRCA of A. carolinensis and A. punctatus has been dated as 49 million years old (Prates et al., 2015). Additional reference genomes representing more lineages within the ancient and highly diverse Anolis radiation will improve inferences of genomic divergence in response to environmental gradients in this system.

82

Table 1.1: Median values and 95% highest posterior density intervals (in millions of years) of log-normally distributed calibration priors applied in dating analyses, based on fossil data.

Settings for calibration prior mean, standard deviation and offset are provided. MRCA = most recent common ancestor.

95% 95% Calibrated Median Prior Prior Prior Fossil HPD HPD node value mean st. dev. offset lower upper Saichangurvel MRCA of 77.39 71.54 105.4 2 0.8 70 davidsoni Pleurodonta Stem of Suzanniwana Basiliscus and 62.39 56.54 90.44 2 0.8 55 sp. Corytophanes Stem of Afairiguana Leiosaurus and 57.39 51.54 85.44 2 0.8 50 avius Urostrophus

83

Table 3.1: Posterior estimates of the time of divergence (T, in millions of years) between lizards in the Atlantic Forest (AF) and eastern Amazonia (EAm), the time of population bottlenecks in the Atlantic Forest, and effective population sizes (Ne, in millions of individuals), based on best- fit scenarios in DIYABC analyses. 95% credibility intervals (CI) are indicated.

Anolis punctatus Lower 95% Upper 95% Median Mean Mode CI CI T EAm-AF divergence 0.545 0.628 0.450 0.194 1.620 T AF population expansion 0.290 0.337 0.240 0.092 0.833 Ne WAm 1.860 1.950 1.610 0.961 3.510 Ne Eam 3.360 3.280 3.770 1.540 4.850 Ne AF bottleneck 0.035 0.040 0.008 0.003 0.094 Ne AF after expansion 3.120 3.160 3.000 1.680 4.650

Anolis ortonii Lower 95% Upper 95% Median Mean Mode CI CI T EAm-AF divergence 0.501 0.565 0.455 0.156 1.240 T AF population expansion 0.276 0.342 0.110 0.032 1.020 Ne WAm 2.700 2.740 2.410 1.430 4.390 Ne EAm 4.530 4.400 4.600 3.040 4.980 Ne AF bottleneck 0.054 0.053 0.064 0.005 0.098 Ne AF after expansion 1.870 2.090 1.190 0.232 4.650

Polychrus marmoratus Lower 95% Upper 95% Median Mean Mode CI CI T EAm-AF divergence 0.740 0.876 0.482 0.136 2.370 Ne EAm-AF ancestor 2.190 2.320 1.400 0.761 4.280 Ne EAm after vicariance 1.290 1.560 0.681 0.191 4.220 Ne AF after vicariance 1.260 1.370 1.170 0.336 3.050

84

Table 4.1: Posterior estimates from G-PhoCS analyses of Anolis ortonii and A. punctatus. Model parameters are the time of coalescence between populations (T, in millions of years), effective population sizes (N, in millions of individuals), and migration rates (M, the proportion of individuals in one population that arrived from another population in each generation) between

Atlantic Forest lizards (af) and their most closely-related Amazonian population (am). Estimates for ancestral populations (anc) and western Amazonian populations of A. punctatus (wamz) are also indicated. CI denotes 95% credibility intervals.

Anolis ortonii Lower 95% Upper 95% Median Mean CI CI T eam-af divergence 0.310 0.312 0.264 0.351 N af 0.177 0.177 0.153 0.201 N eam 6.105 6.116 5.279 6.994 N anc1 0.666 0.666 0.600 0.726 M af -> eam 5.40E-12 1.55E-10 2.42E-14 9.15E-10 M eam -> af 1.04E-11 5.95E-10 2.42E-14 3.70E-09

Anolis punctatus Lower 95% Upper 95% Median Mean CI CI T eam-af divergence 0.231 0.233 0.198 0.269 T anc1-am 0.599 0.600 0.541 0.649 T wamz2-wamz3 0.636 0.638 0.579 0.690 T anc2-anc3 0.967 0.966 0.901 1.025 N af 0.113 0.113 0.096 0.129 N eam 1.032 1.043 0.847 1.240 N wamz1 0.512 0.514 0.458 0.568 N wamz2 2.004 2.004 1.767 2.242 N wamz3 0.340 0.340 0.306 0.373 N anc1 0.416 0.414 0.325 0.496 N anc2 0.395 0.397 0.295 0.498 N anc3 0.335 0.336 0.246 0.430 N anc4 0.633 0.634 0.580 0.687 M af -> eam 1.09E-11 4.96E-10 2.42E-14 2.81E-09 M eam -> af 1.50E-11 1.27E-09 2.42E-14 6.80E-09

85

Table 4.2: Candidate loci recovered by genome-environment association analyses using LFMM for Anolis ortonii and A. punctatus. Only genes that successfully blasted against protein-coding regions of the genome of A. carolinensis and were successfully annotated for gene ontologies are shown. Gene ontology terms (GO names) for biological activities at the molecular (F) and cell or organism level (P) are indicated.

Anolis ortonii Sim. Gene Length #Hits e-Value #GO GO Names (F and P) mean F: transcription cofactor activity; F: protein tyrosine phosphatase activity; F: receptor tyrosine kinase binding; Tyrosine- P: negative regulation of cell phosphatase 91 2 3.81E-11 96 11 proliferation; P: non-receptor lymphangiogenesis; P: type 21 regulation of nucleic acid- templated transcription; P: peptidyl-tyrosine dephosphorylation; P: regulation of protein export from nucleus. MON2 homolog 126 4 2.07E-13 96 1 isoform X1 F: histone binding; Scm-like P: negative regulation of with four transcription, DNA- MBT 91 5 1.17E-06 100 5 templated; domains 1 P: negative regulation of isoform X3 muscle organ development. P: regulation of response to NTPase stimulus; KAP family P: regulation of signaling; P-loop 88 2 1.21E-04 77.5 6 P: negative regulation of domain- apoptotic process; containing 1 P: regulation of cell isoform X1 communication; 86

P: microtubule cytoskeleton organization. F: structural molecule activity; F: alpha-L- arabinofuranosidase activity; P: sensory perception of Mucin-2 89 8 3.62E-13 79.63 7 sound; (predicted) P: L-arabinose metabolic process; P: adult locomotory behavior; P: single-organism developmental process. F: proton-transporting ATPase activity, rotational mechanism; F: ATPase binding; P: ATP hydrolysis coupled proton transport; V-type P: vacuolar acidification; proton P: toxin transport; ATPase 116 108 12 1.21E-07 100 19 P: protein catabolic process in kDa subunit the vacuole; a isoform X1 P: vacuolar proton- transporting V-type ATPase complex assembly; P: ATP synthesis coupled proton transport; P: macroautophagy.

Anolis punctatus Gene Sim. Length #Hits e-Value #GO GO Names (F and P) mean F: zinc ion binding; Mast cell F: metallocarboxypeptidase carboxy- 89 1 1.21E-05 100 4 activity; peptidase A P: proteolysis. Zinc finger F: nucleic acid binding; 97 20 2.78E-11 83.8 2 135 F: metal ion binding. F: serine-type endopeptidase Tissue factor inhibitor activity; pathway 94 1 1.07E-06 87 5 P: negative regulation of inhibitor endopeptidase activity; P: blood coagulation.

87

Mucin-2 89 3 1.75E-08 73 1 (predicted) Zinc finger CCCH domain- 113 2 8.17E-10 92 2 F: metal ion binding. containing 13 isoform X1 P: positive regulation of leukocyte migration; P: wound healing; Melanoma P: negative regulation of cell inhibitory 121 1 2.08E-04 100 8 adhesion; activity 3 P: protein transport; P: exocytosis; P: negative regulation of cell migration. F: calmodulin binding; P: male gonad development; P: spermatogenesis; P: negative regulation of neuron differentiation; P: positive regulation of canonical Wnt signaling pathway; P: positive regulation of neuroblast proliferation; P: cerebral cortex development; Abnormal P: developmental growth; spindle-like P: negative regulation of 89 1 1.33E-10 93 24 microcephaly- asymmetric cell division; associated P: neuronal stem cell population maintenance; P: forebrain neuroblast division; P: spindle localization; P: regulation of meiotic cell cycle; P: neuron migration; P: oogenesis; P: maintenance of centrosome location; P: spindle assembly involved in meiosis. Zinc finger F: nucleic acid binding; 97 20 3.46E-12 91.95 5 709-like F: transcription factor activity, 88

sequence-specific DNA binding; F: metal ion binding; P: regulation of transcription, DNA-templated. Zinc finger 97 20 1.35E-11 79.05 1 F: binding. 420-like F: ubiquitin protein ligase activity; F: ubiquitin protein ligase binding; Ubiquitin- F: ubiquitin conjugating conjugating 108 1 6.48E-14 96 6 enzyme activity; enzyme E2 Z P: protein ubiquitination involved in ubiquitin- dependent protein catabolic process. F: phosphatidylinositol-4- phosphate binding; Oxysterol- F: phosphatidylserine binding - 131 3 1.29E-14 96 6 binding; F: phospholipid related 5 transporter activity; isoform X3 F: cholesterol binding; P: phospholipid transport. F: nucleic acid binding; Zinc finger 25 88 20 6.73E-09 79.7 2 F: metal ion binding. Zinc finger F: nucleic acid binding; 658B isoform 97 20 3.98E-12 90 2 F: metal ion binding. X29 F: protein dimerization activity; P: meiotic chromosome condensation; P: reciprocal meiotic recombination; P: female meiosis sister Stimulated by chromatid cohesion; retinoic acid 111 1 1.38E-04 94 15 P: spermatogenesis; gene 8 P: fertilization; homolog P: regulation of organ growth; P: synapsis; P: meiotic DNA double- strand break formation; P: ovarian follicle development; P: oocyte development; 89

P: cellular response to retinoic acid; P: meiotic DNA replication checkpoint.

90

Figure 1.1: Dewlap coloration patterns of our focal Amazonian anole lizards, Anolis punctatus

(A), A. philopunctatus (B), A. phyllorhinus (C), and A. dissimilis (D).

91

Figure 1.2: Collection sites of new Dactyloa clade samples included in this study. Orange:

Anolis punctatus with orange dewlaps. White: A. punctatus with white dewlaps. Purple: A. philopunctatus. Red: A. phyllorhinus. Green: A. dissimilis. Blue: A. transversalis. Abbreviations of sampled states in Brazil are indicated. 1. Fazenda Experimental Catuaba, Acre (AC): A. dissimilis UFAC0075, UFAC0084, UFAC0089, A. punctatus MTR28593, A. transversalis

MTR28583; 2. Serra do Divisor National Park, AC: A. punctatus MTR28048; 3. Porto Velho,

Rondônia (RO), A. punctatus H1907; 4. Pacaás Novos National Park, RO: A. punctatus

MTR26005; 5. Itapuru, Amazonas (AM): A. punctatus MTR18550; 6. Coari, AM: A. punctatus

MPEG26911; 7. Novo Aripuanã, AM: A. phyllorhinus MTR14041; 8. Colniza, Mato Grosso

(MT): A. phyllorhinus MTR977398, MTR977423, MTR977628, MTR977664, A. punctatus

MTR967985; 9. Jacareacanga, Pará (PA): A. phyllorhinus PJD002; 10. Manaus, AM: A. philopunctatus MTR21474; 11. Presidente Figueiredo, AM: A. philopunctatus Galo132; 12.

Urucará, AM: A. punctatus MPEG29316; 13. Oriximiná, PA: A. punctatus MPEG24758; 14.

Juruti, PA: A. punctatus MPEG28489; 15. Afuá, PA: A. punctatus MPEG29591; 16. Camacan,

Bahia (BA): A. punctatus MTR16109; 17. Linhares, Espírito Santo (ES): A. punctatus

MTR12509.

92

93

Figure 1.3: Dewlap coloration of preserved Anolis punctatus (top) and A. philopunctatus. Both patterns remain promptly discernible even after at 27 years of fixation.

94

Figure 1.4: Coalescent-based phylogeny of the Dactyloa clade of Anolis inferred using

*BEAST. Posterior probabilities = 1 are indicated with an asterisk. Species groups follow

Castañeda and de Queiroz (2013). Picture credits: Julián Velasco (A. calimae).

95

Figure 1.5: Phylogeny of the Dactyloa clade of Anolis based on a concatenated dataset.

Bayesian (MrBayes) and maximum likelihood (Garli) analyses recovered the same topology.

Node values represent posterior probability/bootstrap support values. Posterior probabilities = 1 and bootstrap values = 100 are indicated with an asterisk. Species groups follow Castañeda and de Queiroz (2013). Picture credits: Alejandro Arteaga, Tropical Herping (A. proboscis), Pedro

Peloso (A. transversalis), Renato Recoder (A. punctatus), Bret Whitney (A. phyllorhinus).

96

97

Figure 1.6: Divergence times between pleurodont iguanian lizards, based on five nuclear markers (4,082 base pairs) and inferred using BEAST. Red circles denote calibrated nodes.

Posterior probabilities = 1 are indicated with an asterisk. Bars represent the 95% highest posterior densities (HPD). Picture credits: Pedro Peloso (Anolis trachyderma).

98

Figure 2.1: Coloration in life of Anolis nasofrontalis (A, B) and A. pseudotigrinus (C, D). In A, inset shows the black throat lining of A. nasofrontalis. Photographed specimens are females.

99

Figure 2.2: Phylogenetic relationships and divergence times between species in the Dactyloa clade of Anolis inferred using BEAST. Asterisks denote posterior probabilities > 0.95.

100

Figure 2.3: Head in dorsal and ventral views of Anolis nasofrontalis (A, B) and A. pseudotigrinus (C, D).

101

Figure 2.4: Geographic distribution of confirmed and purported members of the neblininus species series within the Dactyloa clade of Anolis. Anolis bellipeniculus: Cerro Yaví, Amazonas,

Venezuela (Myers and Donelly, 1996); A. calimae: San Antonio, Television Tower Mountain,

Valle del Cauca, Colombia, and Lake Calima Dam, Daríen, Valle del Cauca, Colombia (Ayala et al., 1983); A. carlostoddi: Abacapa-tepui, Chimantá massif, Bolivar, Venezuela (Williams et al.,

1996); A. dissimilis: Fazenda Experimental Catuaba, Senador Guiomard, Acre, Brazil (Melo-

Sampaio et al., 2013; Prates et al., 2015), Parque Estadual Chandless, Manoel Urbano, Acre,

Brazil (Freitas et al., 2013), San Martín, Lower Urubamba Region, Peru (Icochea et al., 2001), and Itahuania, upper Rio Madre de Dios, Madre de Dios, Peru (Williams, 1965); Anolis nasofrontalis and A. pseudotigrinus: Santa Teresa, Espírito Santo, Brazil (this study; Myers and

Carvalho, 1945; Williams and Vanzolini, 1980); A. neblininus: Cerro de la Neblina, Amazonas,

Venezuela (Myers et al., 1993); A. peruensis: Esperanza, Amazonas, Peru (Poe et al., 2015); A. williamsmittermeierorum: Venceremos, San Martin, Peru (Poe and Yañez-Miranda, 2007). For the last two taxa, a single point is depicted on map due to close proximity of collection sites. The inset presents a schematic map of South America around 10-12 mya (based on Lundberg et al.

1998), illustrating the approximate locality of the Chapare buttress, a land bridge that connected the central Andes to the western edge of the Brazilian Shield during the Miocene.

102

103

Figure 3.1: Demographic scenarios tested with coalescent simulations and ABC. A)

Northeastern vicariance: an ancestral population occurring in both eastern Amazonia (EAm) and in the Atlantic Forest (AF) is split due to expansion of open and dry domains in presently dry northeastern South America. B) Northeastern dispersal: a founder population colonizes the AF from EAm, with population expansion following a bottleneck. Scenarios C (southwestern vicariance) and D (southwestern dispersal) are similar to A and B respectively, with the difference that the southwestern (instead of eastern) Amazonia acts as the source of dispersal into the Atlantic Forest. T = time. E) Map showing the three regions considered in this study.

104

Figure 3.2: Sampled sites and coalescent-based phylogeny of Anolis punctatus (a), A. ortonii

(b), and Polychrus marmoratus (c) based on a multi-locus dataset, depicting the relationships between putative independently evolving lineages as inferred based on bGMYC. Colors correspond to the spatial groups used in analyses of alternative demographic scenarios in

DIYABC, as follows: blue, southwestern Amazonia; orange, eastern Amazonia; red; Atlantic

Forest. Brazilian state acronyms on maps are as follows: AC, Acre; AL, Alagoas; AM,

Amazonas; BA, Bahia; ES, Espírito Santo; MA, Maranhão; MG, Minas Gerais; MT, Mato

Grosso; PA, Pará; RJ, Rio de Janeiro; RO, Rondônia; SE, Sergipe; SP, São Paulo; TO,

Tocantins.

105

Figure 4.1: Historical demographic models and parameters estimated using G-PhoCS for Anolis ortonii (top) and A. punctatus. N, effective population sizes; T, time of population coalescence; af, Atlantic Forest population; am, Amazonian population (in A. ortonii); anc, ancestral populations; eam, eastern Amazonian population (in A. punctatus); weam, western Amazonian populations (in A. punctatus).

106

Figure 4.2: Patterns of genetic structure inferred using sNMF for Anolis ortonii (top) and A. punctatus. A population tree of A. punctatus estimated with SNAPP is also indicated.

107

Figure 4.3: Ontology annotations of candidate loci recovered by genome-environment association analyses performed using LFMM, for Anolis ortonii (pies on the left) and A. punctatus. Activities at the level of molecular function (top) and biological process at the level of cell or organism are indicated. For simplicity, pie slices that correspond to minor categories that are not shared by the two anole species are indicated in gray.

108

Appendix 1.1: Newly sampled specimens used in the phylogenetic analyses of Dactyloa in

Chapter 1, including voucher numbers, locality information, and Genbank accession numbers.

Sample BR ND2, Species Locality RAG1 ID State tRNA-Trp Senador Guiomard, UFAC A. dissimilis AC Fazenda Experimental KM598666 KM598698 0075 Catuaba Senador Guiomard, UFAC A. dissimilis AC Fazenda Experimental KM598667 KM598699 0084 Catuaba Senador Guiomard, UFAC A. dissimilis AC Fazenda Experimental KM598668 KM598700 0089 Catuaba Galo Presidente Figueiredo, A. philopunctatus AM KM598669 KM598701 132 UHE Balbina MTR A. philopunctatus AM Manaus KM598670 KM598702 21474 MTR Novo Aripuanã, Rio A. phyllorhinus AM KM598671 KM598703 14041 Roosevelt MTR A. phyllorhinus MT Colniza, Rio Aripuanã KM598672 KM598704 977398 MTR A. phyllorhinus MT Colniza, Rio Aripuanã KM598673 KM598705 977423 MTR A. phyllorhinus MT Colniza, Rio Aripuanã KM598674 KM598706 977628 MTR A. phyllorhinus MT Colniza, Rio Aripuanã KM598675 KM598707 977664 PJD A. phyllorhinus PA Jacareacanga, AHE Jatobá KM598676 KM598708 002 H A. punctatus RO Porto Velho, Caiçara KM598677 KM598709 1907 MPEG Oriximiná, Porto A. punctatus PA KM598678 KM598710 24758 Trombetas MPE A. punctatus AM Coari KM598679 KM598711 26911 MPE A. punctatus PA Juruti, Galiléia KM598680 KM598712 28489 MPE A. punctatus AM Urucará, Marajatuba KM598681 KM598713 29316 MPE A. punctatus PA Afuá, FLONA Charapucu KM598682 KM598714 29591 A. punctatus MTR ES Linhares, Reserva Cia. KM204351 KM598715 109

12509 Vale do Rio Doce MTR A. punctatus BA Camacan, Serra Bonita KM204353 KM598716 16109 MTR Beruri, Itapuru, Lago A. punctatus AM KM598683 KM598717 18550 Chaviana MTR Parque Nacional Serra dos A. punctatus RO KM598684 KM598718 26005 Pacaás Novos, base Jaci MTR Parque Nacional Serra do A. punctatus AC KM598685 KM598719 28048 Divisor, Pé da Serra Senador Guiomard, MTR A. punctatus AC Fazenda Experimental KM598686 KM598720 28593 Catuaba MTR A. punctatus MT Colniza, Rio Aripuanã KM598687 KM598721 967985 Senador Guiomard, MTR A. transversalis AC Fazenda Experimental KM598688 KM598722 28583 Catuaba

110

Appendix 1.2: Pleurodont iguanian species used in divergence time estimation analyses in

Chapter 1, including voucher and Genbank accession numbers.

Species Sample ID BACH1 DNAH3 MKL1 NGFB SNCAIP LSUMZH JF EU EU EU GU Anolis carolinensis 503 805906 402726 390801 437979 432677 UFAC KM KM KM KM Anolis dissimilis - 0084 598732 598741 598689 598723 Anolis PEU KM KM KM KM KM fuscoauratus 488 598736 598745 598753 598693 598727 H KM KM KM KM KM Anolis ortonii 2026 598737 598746 598754 598694 598728 Anolis MTR KM KM KM KM KM phyllorhinus 977664 598733 598742 598750 598690 598724 GN KM KM KM KM KM Anolis punctatus 71 598734 598743 598751 598691 598725 Anolis MTR KM KM KM KM KM transversalis 28583 598735 598744 598752 598692 598726 MTR KM KM KM KM KM Anolis tandai 28575 598738 598747 598755 598695 598729 Anolis MTR KM KM KM KM KM trachyderma 28203 598739 598748 598756 598696 598730 Basiliscus TOMH JF HQ JF JF JF basiliscus 86-271 805907 876403 804430 818336 818388 Chalarodon YPM JF JF JF JF JF madagascariensis 12866 805909 806095 804434 818349 818401 Corytophanes WL JF JF JF JF JF cristatus 8961 805910 806096 804440 818337 818389 Enyalioides KU JF GU JF GU GU laticeps 212627 805913 457910 804446 432714 432675 Leiosaurus LJA JF JF JF JF JF catamarcensis 2047 805916 806098 804454 818344 818396 Morunasaurus MVZ JF HQ JF JF JF annularis 163062 805918 876405 804458 818340 818392 BPN JF JF JF JF JF Plica plica 1052 805928 806104 804479 818358 818410 Polychrus MTR KM KM KM KM KM liogaster 25918 598740 598749 598757 598697 598731 Polychrus ATO JF HQ JF JF JF marmoratus L31 805923 876410 804467 818355 818407 Stenocercus JF HQ JF JF JF TWRStg guentheri 805927 876412 804475 818357 818409 Uranoscodon BPN JF JF JF JF JF 111

superciliosus 805 805930 806106 804481 818359 818411 Urostrophus LSUMZH JF HQ JF JF JF vautieri 13960 805931 876407 804483 818346 818398

112

Appendix 3.1: Sampled specimens used in Chapter 3, including locality information.

BR Species Sample ID Locality State Senador Guiomard: Fazenda A. dissimilis UFAC0084 AC Experimental Catuaba Senador Guiomard: Fazenda A. dissimilis UFAC0087 AC Experimental Catuaba A. Porto Velho: Caicara (left bank of H0805 RO fuscoauratus Madeira River) A. PEU488 BA Ilheus fuscoauratus A. ortonii LG2277 BA Ilheus A. ortonii LSUMZH13904 AC Porto Walter: Rio Jurua Labrea: Rio Ituxi Madeirera A. ortonii LSUMZH14099 AM Scheffer A. ortonii LSUMZH14163 PA Alter de Chao Belterra: Agropecuaria Treviso A. ortonii LSUMZH14264 PA LTDA Belterra: Agropecuaria Treviso A. ortonii LSUMZH14304 PA LTDA Belterra: Agropecuaria Treviso A. ortonii LSUMZH14313 PA LTDA Belterra: Agropecuaria Treviso A. ortonii LSUMZH14406 PA LTDA Belterra: Agropecuaria Treviso A. ortonii LSUMZH14444 PA LTDA A. ortonii MPEG25698 PA Portel: Floresta Nacional Caxiuana A. ortonii MPEG25699 PA Portel: Floresta Nacional Caxiuana A. ortonii MPEG27088 PA Juruti: Mutum Belem: Campus de Pesquisa do A. ortonii MPEG27688 PA MPEG A. ortonii MPEG28256 PA Juruti: road to Barroso A. ortonii MPEG29325 AM Urucara: Marajatuba A. ortonii MPEG30019 PA Itaituba (left bank) Senador Guiomard: Fazenda A. ortonii UFAC0085 AC Experimental Catuaba Jussari: Reserva Particular do A. ortonii LG1783 BA Patrimonio Natural Serra do Teimoso Linhares: Reserva da Companhia A. ortonii MTR12160 ES Vale do Rio Doce Linhares: Floresta Nacional de A. ortonii MTR12291 ES Goytacazes 113

Linhares: Floresta Nacional de A. ortonii MTR12329 ES Goytacazes A. ortonii MTR18000 SE Itabaiana Belmonte: Barrolandia E. E. A. ortonii MTR34278 BA Gregorio Bondar CEPLAC Belmonte: Barrolandia E. E. A. ortonii MTR34436 BA Gregorio Bondar CEPLAC A. ortonii BM028 PA Vitoria do Xingu: UHE Belo Monte A. ortonii BM565 PA Anapu Porto Velho: Abuna (right bank of A. ortonii H1672 RO Madeira River) Porto Velho: Mutum (right bank of A. ortonii H2026 RO Madeira River) Porto Velho: Abuna (left bank of A. ortonii H2594 RO Madeira River) A. ortonii H4251 RO Porto Velho: Mutum A. ortonii H5121 RO Porto Velho: Abuna A. ortonii MTR28130 AC Parque Nacional Serra do Divisor A. ortonii MTR977671 MT Aripuana: Rio Aripuana A. ortonii MTR977673 MT Aripuana: Rio Aripuana A. phyllorhinus MTR14041 AM Novo Aripuana: Rio Roosevelt A. phyllorhinus MTR977664 MT Colniza: Rio Aripuana A. punctatus CTMZ03457 RO Montenegro A. punctatus CTMZ03491 RO Montenegro A. punctatus CTMZ03502 RO Montenegro A. punctatus CTMZ03507 RO Montenegro A. punctatus CTMZ03926 AL Campo Alegre Labrea: Rio Ituxi Madeirera A. punctatus LSUMZH14100 AM Scheffer Labrea: Rio Ituxi Madeirera A. punctatus LSUMZH14101 AM Scheffer Labrea: Rio Ituxi Madeirera A. punctatus LSUMZH14102 AM Scheffer Belterra: Agropecuaria Treviso A. punctatus LSUMZH14225 PA LTDA Belterra: Agropecuaria Treviso A. punctatus LSUMZH14336 PA LTDA Belterra: Agropecuaria Treviso A. punctatus LSUMZH14453 PA LTDA Nova Mamore: Parque Estadual A. punctatus LSUMZH15475 RO Guajara Mirim Nova Mamore: Parque Estadual A. punctatus LSUMZH15476 RO Guajara Mirim Nova Mamore: Parque Estadual A. punctatus LSUMZH17842 RO Guajara Mirim Rio Formoso 114

A. punctatus MPEG20750 PA Oriximina A. punctatus MPEG20846 PA Juruti: Mutum Guajara Mirim: Parque Nacional A. punctatus MPEG21348 RO Serra da Cutia Guajara Mirim: Parque Nacional A. punctatus MPEG21355 RO Serra da Cutia A. punctatus MPEG21863 PA Juruti Santa Barbara do Para: Parque A. punctatus MPEG22415 PA Estadual de Gunma A. punctatus MPEG22857 PA Almeirim A. punctatus MPEG24750 PA Oriximina: Porto Trombetas A. punctatus MPEG24758 PA Oriximina: Porto Trombetas A. punctatus MPEG25131 PA Juruti: Galileia A. punctatus MPEG26101 PA Portel: Floresta Nacional Caxiuana A. punctatus MPEG26102 PA Portel: Floresta Nacional Caxiuana A. punctatus MPEG27817 PA Juruti: Capiranga A. punctatus MPEG28489 PA Juruti: Galileia A. punctatus MPEG29316 AM Urucara: Marajatuba A. punctatus MPEG29453 PA Itaituba: Sao Luis (right bank) Afua: Floresta Nacional Charapucu A. punctatus MPEG29591 PA Rio Preto A. punctatus MPEG29943 PA Ruropolis A. punctatus MPEG29944 PA Ruropolis A. punctatus MPEG29983 PA Almeirim: Monte Dourado A. punctatus IBSPCRIB0361 SP Bertioga Cariacica: Reserva Biologica Duas A. punctatus JFT459 ES Bocas Cariacica: Reserva Biologica Duas A. punctatus JFT467 ES Bocas A. punctatus JFT773 ES Santa Teresa: Recanto da Preguica A. punctatus LG1299 BA Una A. punctatus LG1660 BA Una A. punctatus MTR (1468) ES Sao Jose do Calcado: UHE Rosal A. punctatus MTR (1478) RJ Rio de Janeiro: PARNA da Tijuca Jussari: Reserva Particular do A. punctatus MTR05978 BA Patrimonio Natural Serra do Teimoso Linhares: Reserva da Companhia A. punctatus MTR12103 ES Vale do Rio Doce Linhares: Floresta Nacional de A. punctatus MTR12131 ES Goytacazes Linhares: Reserva da Companhia A. punctatus MTR12161 ES Vale do Rio Doce Linhares: Floresta Nacional de A. punctatus MTR12238 ES Goytacazes 115

Linhares: Reserva da Companhia A. punctatus MTR12509 ES Vale do Rio Doce Linhares: Reserva da Companhia A. punctatus MTR12510 ES Vale do Rio Doce Linhares: Reserva da Companhia A. punctatus MTR12511 ES Vale do Rio Doce Linhares: Reserva da Companhia A. punctatus MTR13678 ES Vale do Rio Doce Cachoeiras de Macacu: Parque A. punctatus MTR15267 RJ Estadual dos Tres Picos A. punctatus MTR16109 BA Camacan: Reserva Serra Bonita A. punctatus MTR16124 BA Camacan: Reserva Serra Bonita Marlieria: Parque Estadual do Rio A. punctatus MTR17744 MG Doce Sooretama: Reserva Bioloogica A. punctatus MTR21545 ES Sooretama Bemonte: Santa Maria Eterna Haras A. punctatus MTR33057 BA Ranchito A. punctatus MTR33102 BA Belmonte: Santa Maria Eterna Bemonte: Santa Maria Eterna Haras A. punctatus MTR33120 BA Ranchito Bemonte: Santa Maria Eterna Haras A. punctatus MTR33121 BA Ranchito A. punctatus MTR33197 BA Canavieiras: Fazenda Cotovelo A. punctatus MTR33210 BA Belmonte: Fazenda Tuiuti A. punctatus MTR33211 BA Belmonte: Fazenda Tuiuti A. punctatus MTR33212 BA Belmonte: Fazenda Tuiuti Belmonte: Barrolandia E. E. A. punctatus MTR34109 BA Gregorio Bondar CEPLAC Belmonte: Barrolandia E. E. A. punctatus MTR34157 BA Gregorio Bondar CEPLAC Belmonte: Barrolandia E. E. A. punctatus MTR34208 BA Gregorio Bondar CEPLAC A. punctatus MTR34227 BA Itapebi: Fazenda Boa Sorte A. punctatus MTR34230 BA Itapebi: Fazenda Boa Sorte A. punctatus MTR34337 BA Itapebi: Fazenda Boa Sorte A. punctatus MTR34414 BA Porto Seguro: Estacao Veracel Belmonte: Barrolandia E. E. A. punctatus MTR34453 BA Gregorio Bondar CEPLAC Santa Tereza: Reserva Biologica A. punctatus MTR34692 ES Augusto Ruschi Santa Tereza: Reserva Biologica A. punctatus MTR34693 ES Augusto Ruschi Santa Tereza: Reserva Biologica A. punctatus MTR34722 ES Augusto Ruschi 116

A. punctatus MTR35172 BA Parque Estadual Serra do Conduru A. punctatus MUFAL9635 AL Maceio A. punctatus BM288 PA Vitoria do Xingu: UHE Belo Monte A. punctatus BM476 PA Vitoria do Xingu: UHE Belo Monte A. punctatus GN71 TM372 MT Guaranta do Norte A. punctatus GN72 TM373 MT Guaranta do Norte A. punctatus MTR976158 MT Claudia A. punctatus MTR976723 MT Gaucha do Norte A. punctatus MTR978312 MT Vila Rica Porto Velho: Caicara (right bank of A. punctatus H0737 RO Madeira River) Porto Velho: Mutum (right bank of A. punctatus H0877 RO Madeira River) Porto Velho: Abuna (left bank of A. punctatus H1100 RO Madeira River) Porto Velho: Caicara (left bank of A. punctatus H1907 RO Madeira River) Porto Velho: Caicara (right bank of A. punctatus H1911 RO Madeira River) Porto Velho: Abuna (left bank of A. punctatus H2287 RO Madeira River) A. punctatus H2546 RO Porto Velho: Abuna A. punctatus H3039 RO Porto Velho: Mutum A. punctatus H3306 RO Porto Velho: Caicara A. punctatus H3972 RO Porto Velho: Caicara Parque Nacional Serra dos Pacaas A. punctatus MTR25584 RO Novos Base Candeias Parque Nacional Serra dos Pacaas A. punctatus MTR25585 RO Novos Base Candeias Parque Nacional Serra dos Pacaas A. punctatus MTR25604 RO Novos Base Candeias Parque Nacional Serra dos Pacaas A. punctatus MTR25618 RO Novos Base Candeias Parque Nacional Serra dos Pacaas A. punctatus MTR25673 RO Novos Base Candeias Parque Nacional Serra dos Pacaas A. punctatus MTR25931 RO Novos Base Candeias Parque Nacional Serra dos Pacaas A. punctatus MTR26005 RO Novos Base Jaci A. punctatus MTR967985 MT Colniza: Rio Aripuana A. scypheus MTR33729 AM Rio Japura A. scypheus MTR33800 AM Rio Japura A. tandai MTR28259 AC Parque Nacional Serra do Divisor Senador Guiomard: Fazenda A. tandai MTR28575 AC Experimental Catuaba 117

A. MTR28049 AC Parque Nacional Serra do Divisor trachyderma A. MTR28203 AC Parque Nacional Serra do Divisor trachyderma A. MTR18743 AM Beruri: Itapuru Lago Chaviana transversalis A. Senador Guiomard: Fazenda MTR28583 AC transversalis Experimental Catuaba Parque Nacional Serra dos Pacaas P. liogaster MTR25918 RO Novos Base Candeias Senador Guiomard: Fazenda P. liogaster MTR28582 AC Experimental Catuaba P. marmoratus MPEG21807 PA Ourem: Fazenda Urubu Oriximina Porto Trombetas. Plato P. marmoratus MPEG24799 PA Bacaba Belem: Campus de Pesquisa do P. marmoratus MPEG25076 PA MPEG Guama P. marmoratus MPEG27038 PA Barcarena: Vila dos Cabanos P. marmoratus MPEG27039 PA Barcarena: Vila dos Cabanos P. marmoratus LG1001 BA Porto Seguro Linhares: Reserva da Companhia P. marmoratus MTR12079 ES Vale do Rio Doce Linhares: Reserva da Companhia P. marmoratus MTR12421 ES Vale do Rio Doce P. marmoratus MTR15531 AL Murici: Usina Ibitinga P. marmoratus MTR16125 BA Camacan: Reserva Serra Bonita Jequitinhonha: Reserva Biologica P. marmoratus MTR17267 MG da Mata Escura Jequitinhonha: Reserva Biologica P. marmoratus MTR17429 MG da Mata Escura P. marmoratus MTR33099 BA Belmonte: Santa Maria Eterna P. marmoratus MTR33100 BA Belmonte: Santa Maria Eterna P. marmoratus MTR33101 BA Belmonte: Santa Maria Eterna Canavieiras: Associacao Irmaos P. marmoratus MTR33196 BA Unidos P. marmoratus MTR33209 BA Belmonte: Fazenda Tuiuti Belmonte: Barrolandia E. E. P. marmoratus MTR34287 BA Gregorio Bondar CEPLAC Belmonte: Barrolandia E. E. P. marmoratus MTR34440 BA Gregorio Bondar CEPLAC Belmonte: Barrolandia E. E. P. marmoratus MTR34441 BA Gregorio Bondar CEPLAC P. marmoratus MTR34560 BA Itapebi: Fazenda Boa Sorte P. marmoratus ESTR42 MA Carolina P. marmoratus MTR07133 TO Guarai 118

Appendix 3.2: Mitochondrial gene tree of Anolis punctatus (A), A. ortonii (B), and Polychrus marmoratus (C) estimated in Chapter 3.

A

119

B

C

120

Appendix 3.3: Plots of the PC analyses and of the posterior distributions of model parameters from DIYABC analyses in Chapter 3.

121

Appendix 3.4: Posterior estimates of substitution rates for the sampled loci (in substitutions per

DNA site/year), from DIYABC analyses in Chapter 3.

122

Appendix 3.5: Plots of the validation analyses based on pseudo-observed datasets in dpp- msbayes analyses in Chapter 3.

123

Appendix 4.1: Sampled specimens used in Chapter 4, including locality information.

Species Sample ID Country State Locality Latitude Longitude A. ortonii H2594 BR RO Abunã -9.6341 -65.4399 LSUMZH A. ortonii BR PA Alter de Chão -2.5027 -54.9718 14163 A. ortonii BM565 BR PA Anapú -3.4051 -51.4222 Belém, MPEG Campus de A. ortonii BR PA -1.4510 -48.4453 27688 Pesquisa do MPEG Belterra: LSUMZH A. ortonii BR PA Agropecuária -3.1456 -54.8396 14304 Treviso LTDA MTR CEPLAC, A. ortonii BR BA -16.0944 -39.2748 34278 Barrolandia MTR CEPLAC, A. ortonii BR BA -16.0944 -39.2748 34436 Barrolandia Fazenda UFAC A. ortonii BR AC Experimental -10.0679 -67.6063 0085 Catuaba Floresta MTR Nacional de A. ortonii BR ES -19.4483 -40.0804 12291 Goytacazes, Linhares MTR A. ortonii BR BA Ilheus -14.7972 -39.0344 09910 MPEG Itaituba, São A. ortonii BR PA -4.4725 -56.2461 29465 Luis Lábrea: Rio LSUMZH Ituxi A. ortonii BR AM -8.3464 -65.7161 14099 Madeirera Scheffer Lago MTR Chaviana, A. ortonii BR AM -4.3072 -61.8155 18907 Itapuru, Rio Purus MTR Moiobamba, A. ortonii BR AM -4.7279 -62.1321 19241 Rio Purus A. ortonii H2026 BR RO Mutum -9.5981 -65.0478 Portel, MPEG FLONA A. ortonii BR PA -1.9936 -51.6153 25699 Caxiuanã, Plote PPBIO 124

LSUMZH Porto Walter: A. ortonii BR AC -8.2509 -72.7677 13904 Rio Juruá Reserva da Cia MTR A. ortonii BR ES Vale do Rio -19.1517 -40.0644 12160 Doce, Linhares Reserva Faunistica LSUMZH A. ortonii EC SU Cuyabeno, -0.2172 -75.8069 12993 Neotropic Turis MTR A. ortonii BR MT Rio Aripuanã -10.1758 -59.4500 977671 MTR Serra do A. ortonii BR AC -7.4388 -73.6582 28130 Divisor UHE Belo Monte, Vitória A. ortonii BM028 BR RO -3.3228 -51.8448 do Xingu, Rio Xingu MPEG Urucará, A. ortonii BR AM -2.3867 -57.6395 29325 Marajatuba A. H2546 BR RO Abunã -9.5979 -65.3576 punctatus Almeirim, A. MPEG BR PA Monte -0.6287 -52.7390 punctatus 29983 Dourado Belterra: A. LSUMZH BR PA Agropecuária -3.1354 -54.8416 punctatus 14336 Treviso LTDA A. IBSPCRIB BR SP Bertioga -23.7564 -46.0247 punctatus 0361 A. MTR Cachoeira do BR RJ -22.4354 -42.6216 punctatus 15267 Macacu A. H1907 BR RO Caiçara -9.4459 -64.8247 punctatus A. H1911 BR RO Caiçara -9.4358 -64.8203 punctatus A. ICST BR AL Campo Alegre -9.8092 -36.2838 punctatus 764 Divisa A. UNIBAN BR RO Montenegro e -10.2987 -63.2375 punctatus 1670 Cacaulândia Estacel A. MTR BR BA Veracel, Porto -16.4420 -39.0651 punctatus 34414 Seguro A. MTR BR AC Fazenda -10.0679 -67.6063 125

punctatus 28593 Experimental Catuaba A. MTR Fazendo Boa BR BA -15.8958 -39.6611 punctatus 34227 sorte, Itapebi Fl. Nac. de A. MTR BR ES Goyatacazes, -19.4483 -40.0804 punctatus 12338 Linhares A. MTR Gaúcha do BR MT -13.1833 -53.2564 punctatus 976723 Norte Guajará A. MPEG BR RO Mirim: PN -11.6996 -64.3915 punctatus 21348 Serra da Cutia A. Guarantã do GN71 BR MT -9.7307 -54.6064 punctatus Norte A. MPEG BR PA Juruti -2.6096 -56.1842 punctatus 20846 A. MPEG BR PA Juruti, Galiléia -2.5475 -56.2256 punctatus 28489 Lábrea: Rio A. LSUMZH Ituxi BR AM -8.3464 -65.7161 punctatus 14100 Madeirera Scheffer Lago A. MTR Chaviana, BR AM -4.3074 -61.8135 punctatus 18550 Itapuru, Rio Purus A. MPEG BR AM Lindóia -2.9096 -59.0495 punctatus 29314 A. MUFAL BR AL Maceio -9.5073 -35.7099 punctatus 9635 A. MTR BR AM Manaus -2.9979 -59.9494 punctatus 21474 Nova Mamoré: A. LSUMZH BR RO PE Guajará -10.3167 -64.5500 punctatus 15476 Mirim Oriximiná, A. MPEG Porto BR PA -1.7731 -56.3706 punctatus 24758 Trombetas, Platô Bacaba A. MTR P. N. Serra dos BR RO -10.7868 -63.6281 punctatus 25584 Pacaás Novos A. MTR BR RR Pacaraima 4.4621 -61.1451 punctatus 20798 A. MTR Parque BR MG -19.6411 -42.5333 punctatus 17744 Estadual do 126

Rio Doce, Marliéria Portel, A. MPEG FLONA BR PA -1.9969 -51.6418 punctatus 26102 Caxiuanã, Plote PPBIO A. LSUMZH Porto Walter: BR AC -8.2509 -72.7677 punctatus 13910 Rio Juruá Rebio Duas A. JFT459 BR ES Bocas, -20.2811 -40.5219 punctatus Cariacica Recanto da A. JFT773 BR ES Preguica, Santa -19.9600 -40.5200 punctatus Teresa Reserva A. MTR BR ES Biológica de -19.0556 -40.1473 punctatus 21545 Sooretama Reserva A. MTR BR ES CVRD, -19.0564 -39.8942 punctatus 12511 Linhares Reserva A. LSUMZH EC SU Faunistica -0.2172 -75.8069 punctatus 12577 Cuyabeno Reserva Faunistica A. LSUMZH EC SU Cuyabeno, -0.2172 -75.8069 punctatus 12751 Neotropic Turis Rio de Janeiro, A. MTR BR RJ Floresta da -22.9712 -43.2547 punctatus (1478) Tijuca A. MPEG BR PA Rurópolis -4.1144 -55.6771 punctatus 29943 Santa Bárbara do Pará, A. MPEG BR PA Parque -1.2169 -48.2948 punctatus 22415 Ecológico Gunma A. MTR Serra do BR AC -7.4424 -73.6575 punctatus 28048 Divisor Serra do A. MTR BR AC Divisor, face -7.5016 -73.7214 punctatus 28401 Oeste A. MTR Serra do BR BA -15.1547 -39.5233 punctatus 05978 Teimoso, 127

Jussari A. PJD BR PA UHE Jatobá -5.8174 -57.3956 punctatus 409 A. MTR BR ES UHE Rosal -20.9031 -41.7192 punctatus (1468) A. LG1299 BR BA Una -15.2696 -39.0694 punctatus A. MTR BR MT Vila Rica -9.9000 -51.2000 punctatus 978312 A. Vitória do BM288 BR PA -3.3228 -51.8448 punctatus Xingu

128

Bibliography

Akaike, H., 1974. A new look at the statistical model identification. IEEE Trans. Autom. Control 19, 716–723. Akashi, H.D., Cádiz Díaz, A., Shigenobu, S., Makino, T., Kawata, M., 2016. Differentially expressed genes associated with adaptation to different thermal environments in three sympatric Cuban Anolis lizards. Mol. Ecol., 25, 2273-2285. Alföldi, J., Di Palma, F., Grabherr, M., Williams, C., Kong, L., Mauceli, E., Russell, P., Lowe, C.B., Glor, R., Jaffe, J.D., Ray, D.A., 2011. The genome of the green anole lizard and a comparative analysis with birds and mammals. Nature 477, 587-591. Amaral, A., 1933. Estudos sobre lacertilios neotropicos. I. Novos gêneros e espécies de lagartos do Brasil. Mem. Inst. Butantan 7, 51–74. Amaro, R.C., Rodrigues, M.T., Yonenaga-Yassuda, Y., Carnaval, A.C., 2012. Demographic processes in the montane Atlantic rainforest: molecular and cytogenetic evidence from the endemic frog Proceratophrys boiei. Mol. Phylogenet. Evol. 62, 880–888. Angulo, A., Icochea, J., 2010. Cryptic species complexes, widespread species and conservation: lessons from Amazonian frogs of the Leptodactylus marmoratus group (Anura: ). Syst. Biodiv. 8, 357–370. Antonelli, A., Quijada-Mascareñas, A., Crawford, A.J., Bates, J.M., Velazco, P.M., Wüster, W., 2010. Molecular studies and phylogeography of Amazonian tetrapods and their relation to geological and climatic models, in: Hoorn, C., Wesselingh, F. (Eds.), Amazonia, Landscape and Species Evolution: A Look into the Past. Wiley-Blackwell, Hoboken, pp. 386-404. Auler, A.S., Wang, X., Edwards, R.L., Cheng, H., Cristalli, P.S., Smart, P.L., Richards, D.A., 2004. Quaternary ecological and geomorphic changes associated with rainfall events in presently semi-arid northeastern Brazil. J. Quat. Sci. 19, 693-701. Avila-Pires, T.C., 1995. Lizards of Brazilian Amazonia (Reptilia: ). Zool. Verh. 299, 1- 706. Ayala, S.C., Harris, D., Williams, E.E., 1983. New or problematic Anolis from Colombia. I. Anolis calimae, new species, from the cloud forest of western Colombia. Breviora 475, 1–11. Ayala-Varela, F.P., Troya-Rodríguez, D., Talero-Rodríguez, X., Torres-Carvajal, O., 2014. A new Andean anole species of the Dactyloa clade (Squamata: Iguanidae) from western . Amphib. Reptile Conserv. 8, 8-24. Barros, F.C., de Carvalho, J.E., Abe, A.S., Kohlsdorf, T., 2010. Fight versus flight: the interaction of temperature and body size determines antipredator behaviour in tegu lizards. Animal Behaviour, 79, 83-88. Barros, T., Williams, E.E., Viloria, A.L., 1996. The genus Phenacosaurus (Squamata: Iguania) in western Venezuela: Phenacosaurus tetarii new species, Phenacosaurus euskalerriari, new species, and Phenacosaurus nicefori Dunn, 1944. Breviora 504, 1-30. Batalha-Filho, H., Fjeldsa, J., Fabre, P., Miyaki, C.M., 2013. Connections between the Atlantic and the Amazonian forest avifaunas represent distinct historical events. J. Ornithol. 154, 41– 50. Beaumont, M.A., 2010. Approximate Bayesian computation in evolution and ecology. Annu. Rev. Ecol. Evol. Syst. 41. Berthold, A.A., 1846. Mittheilungen über das zoologische Museum zu Göttingen. I. Verzeichnisss der aufgestellten Reptilien. Nachr. G. A. Univ. K. Gesell. Wiss. Göttingen 8- 10, 1-28. 129

Bigarella, J.J., Andrade-Lima, D.D., Riehs, P.J., 1975. Considerações a respeito das mudanças paleoambientais na distribuição de algumas espécies vegetais e animais no Brasil. An. Acad. Bras. Cienc. 47. Bonnaterre, P.J., 1789. Tableau encyclopédique et méthodique des trois règnes de la nature: Erpétologie. Panckoucke, Paris. Bouckaert, R., Heled, J., Kühnert, D., Vaughan, T., Wu, C.H., Xie, D., Suchard, M.A., Rambaut, A., Drummond, A.J., 2014. BEAST 2: a software platform for Bayesian evolutionary analysis. PLoS Comput. Biol. 10, e1003537. Brandley, M.C., Wang, Y., Guo, X., de Oca, A.N.M., Fería-Ortíz, M., Hikida, T., Ota, H., 2011. Accommodating heterogenous rates of evolution in molecular divergence dating methods: an example using intercontinental dispersal of Plestiodon (Eumeces) lizards. Syst. Biol. 60, 3- 15. Bridle, J.R., Vines, T.H., 2007. Limits to evolution at range margins: when and why does adaptation fail?. Trends Ecol. Evol., 22, 140-147. Brown, J.L., Twomey, E., 2009. Complicated histories: three new species of poison frogs of the genus Ameerega (Anura: Dendrobatidae) from north-central Peru. Zootaxa 2049, 1-38. Brown, R.M., Siler, C.D., Oliveros, C.H., Esselstyn, J.A., Diesmos, A.C., Hosner, P.A., Linkem, C.W., Barley, A.J., Oaks, J.R., Sanguila, M.B., Welton, L.J., 2013. Evolutionary processes of diversification in a model island archipelago. Annu. Rev. Ecol. Evol. Syst. 44, 411-435. Bryant, D., Bouckaert, R., Felsenstein, J., Rosenberg, N.A., RoyChoudhury, A., 2012. Inferring species trees directly from biallelic genetic markers: bypassing gene trees in a full coalescent analysis. Mol. Biol. Evol., 29, 1917-1932. Buskirk, J.V., 2009. Natural variation in morphology of larval : Phenotypic plasticity in nature?. Ecol. Monog., 79, 681-705. Camacho, C., Coulouris, G., Avagyan, V., Ma, N., Papadopoulos, J., Bealer, K., Madden, T.L., 2009. BLAST+: architecture and applications. BMC Bioinformatics, 10, 421. Campbell-Staton, S.C., Edwards, S.V. and Losos, J.B., 2016. Climate-mediated adaptation after mainland colonization of an ancestrally subtropical island lizard, Anolis carolinensis. J. Evol. Biol. 29, 2168-2180. Carnaval, A.C., Hickerson, M.J., Haddad, C.F., Rodrigues, M.T. Moritz, C., 2009. Stability predicts genetic diversity in the Brazilian Atlantic forest hotspot. Science 323, 785-789. Carnaval, A.C., Waltari, E., Rodrigues, M.T., Rosauer, D., van der Wal, J., Damasceno, R., Prates, I., Strangas, M., Spanos, Z., Rivera, D., Pie, M.R., Firkowski, C.R., Bornschein, M.R., Ribeiro, L.F., Moritz, C., 2014. Prediction of phylogeographic endemism in an environmentally complex biome. Proc. R. Soc. B 281, 1461. Castañeda, M.D.R., Sherratt, E., Losos, J.B., 2014. The Mexican amber anole, Anolis electrum, within a phylogenetic context: implications for the origins of Caribbean anoles. Zool. J. Linn. Soc. 172, 133-144. Castañeda, M.D.R., de Queiroz, K., 2011. Phylogenetic relationships of the Dactyloa clade of Anolis lizards based on nuclear and mitochondrial DNA sequence data. Mol. Phyl. Evol. 61, 784-800. Castañeda, M.D.R., de Queiroz, K., 2013. Phylogeny of the Dactyloa clade of Anolis lizards: New insights from combining morphological and molecular data. Bull. Mus. Comp. Zool. 160, 345–398. Castroviejo-Fisher, S., Guayasamin, J.M., Gonzalez-Voyer, A., Vilà, C., 2014. Neotropical diversification seen through glassfrogs. J. Biogeogr. 41, 66–80. 130

Castroviejo-Fisher, S., Padial, J.M., De La Riva, I., Pombal Jr., J.P., Da Silva, H.R., Rojas- Runjaic, F.J., Medina-Mendez, E., Frost, D.R., 2015. Phylogenetic systematics of egg- brooding frogs (Anura: Hemiphractidae) and the evolution of direct development. Zootaxa 4004, 1-75. Cheng, H., Sinha, A., Cruz, F.W., Wang, X., Edwards, R.L., d’Horta, F.M., Ribas, C.C., Vuille, M., Stott, L.D., Auler, A.S., 2013. Climate change patterns in Amazonia and biodiversity. Nature Comm. 4, 1411. Conesa, A., Götz, S., García-Gómez, J.M., Terol, J., Talón, M., Robles, M., 2005. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics 21, 3674-3676. Conrad, J.L., Norell, M.A., 2007. A complete Late Cretaceous iguanian (Squamata, Reptilia) from the Gobi and identification of a new iguanian clade. Am. Mus. Novit. 3584, 1-47. Conrad, J.L., Rieppel, O., Grande, L., 2007. A Green River (Eocene) polychrotid (Squamata: Reptilia) and a re-examination of iguanian systematics. J. Paleontol. 81, 1365-1373. Cope, E.D., 1868. An examination of the Reptilia and Batrachia obtained by the Orton Expedition to Equador and the Upper Amazon, with notes on other species. Proc. Acad. Nat. Sci. Philadelphia 20, 96-140. Cope, E.D., 1876. Report on the brought by professor James Orton from the middle and upper Amazon and western Peru. Proc. Acad. Nat. Sci. Philadelphia 2, 159–183. Cornuet, J.M., Pudlo, P., Veyssier, J., Dehne-Garcia, A., Gautier, M., Leblois, R., Marin, J.M., Estoup, A., 2014. DIYABC v2. 0: a software to make approximate Bayesian computation inferences about population history using single nucleotide polymorphism, DNA sequence and microsatellite data. Bioinformatics btt763. Costa, L.P. 2003. The historical bridge between the Amazon and the Atlantic Forest of Brazil: a study of molecular phylogeography with small mammals. J. Biogeog. 30, 71-86. Crowley, S.R., 1985. Thermal sensitivity of sprint-running in the lizard Sceloporus undulatus: support for a conservative view of thermal physiology. Oecologia 66, 219-225. Csilléry, K., Blum, M.G., Gaggiotti, O.E., François, O., 2010. Approximate Bayesian computation (ABC) in practice. Trends Ecol. Evol. 25, 410-418. D'Angiolella, A.B., Gamble, T., Avila-Pires, T.C.S., Colli, G.R., Noonan, B.P., 2011. Anolis chrysolepis Duméril and Bibron, 1837 (Squamata: Iguanidae), revisited: molecular phylogeny and taxonomy of the Anolis chrysolepis species group. Bull. Mus. Comp. Zool. 160, 35-63. Danecek, P., Auton, A., Abecasis, G., Albers, C.A., Banks, E., DePristo, M.A., Handsaker, R.E., Lunter, G., Marth, G.T., Sherry, S.T. and McVean, G., 2011. The variant call format and VCFtools. Bioinformatics, 27, 2156-2158. Darriba, D., Taboada, G.L., Doallo, R., Posada, D., 2012. jModelTest 2: more models, new heuristics and parallel computing. Nature Methods 9, 772-772. Daudin, F.M., 1802. Histoire Naturelle, Générale et Particulière des Reptiles, Vol. 4. F. Dufart, Paris. Dennenmoser, S., Vamosi, S.M., Nolte, A.W., Rogers, S.M., 2017. Adaptive genomic divergence under high gene flow between freshwater and brackish‐water ecotypes of prickly sculpin (Cottus asper) revealed by Pool-Seq. Mol. Ecol., 26, 25-42. de Oliveira, P.E., Barreto, A.M.F., Suguio, K., 1999. Late Pleistocene/Holocene climatic and vegetational history of the Brazilian caatinga: the fossil dunes of the middle Sao Francisco River. Palaeogeog. Palaeoclim. Palaeoecol. 152, 319-337. 131

de Queiroz, K., Chu, L.R., Losos, J.B., 1998. A second Anolis lizard in Dominican amber and the systematics and ecological morphology of Dominican amber anoles. Am. Mus. Nov. 3249, 1-23. de Vivo, M., Carmignotto, A.P., 2004. Holocene vegetation change and the mammal faunas of South America and Africa. J. Biogeog. 31, 943-957. Drummond, A.J., Ho, S.Y., Phillips, M.J., Rambaut, A., 2006. Relaxed phylogenetics and dating with confidence. PLoS Biol. 4, e88. Drummond, A.J., Suchard, M.A., Xie, D., Rambaut, A., 2012. Bayesian phylogenetics with BEAUti and the BEAST 1.7. Mol. Biol. Evol. 29, 1969–1973. Duméril A.M.C., Bibron G., 1837. Erpétologie Générale ou Histoire Naturelle Complete des Reptiles, vol. 4. Librairie Encyclopédique de Roret, Paris. Duméril, A.M.C., Duméril, A.H.A., 1851. Catalogue méthodique de la collection des reptiles du Muséum d'Histoire Naturelle de Paris. Gide et Baudry Roret, Paris. Edwards, S.V., 2009. Is a new and general theory of molecular systematics emerging?. Evolution 63, 1-19. Elshire, R.J., Glaubitz, J.C., Sun, Q., Poland, J.A., Kawamoto, K., Buckler, E.S., Mitchell, S.E., 2011. A robust, simple genotyping-by-sequencing (GBS) approach for high diversity species. PloS One 6, e19379. Fagundes, N.J., Ray, N., Beaumont, M., Neuenschwander, S., Salzano, F.M., Bonatto, S.L., Excoffier, L., 2007. Statistical evaluation of alternative models of human evolution. Proc. Nat. Acad. Sci. 104, 17614-17619. Fallon, S.M., Bermingham, E., Ricklefs, R.E., 2005. Host specialization and geographic localization of avian malaria parasites: a regional analysis in the Lesser Antilles. Am. Nat. 165, 466-480. Fiaschi, P., Pirani, J.R., 2009. Review of plant biogeographic studies in Brazil. J. Syst. Evol. 47, 477-496. Flot, J.F., 2010. SeqPHASE: a web tool for interconverting PHASE input/output files and FASTA sequence alignments. Mol. Ecol. Resour. 10, 162-166. Fouquet, A., Loebmann, D., Castroviejo-Fisher, S., Padial, J.M., Orrico, V.G.D., Lyra, M.L., Roberto, I.J., Kok, P.J.R., Haddad, C.F.B., Rodrigues, M.T., 2012b. From Amazonia to the Atlantic forest: molecular phylogeny of Phyzelaphryninae frogs reveals unexpected diversity and a striking biogeographic pattern emphasizing conservation challenges. Mol. Phylogenet. Evol. 65, 547–561. Fouquet, A., Recoder, R.S., Teixeira Jr., M.T., Cassimiro, J., Amaro, R.C., Camacho-Guerrero, A., Damasceno, R., Carnaval, A.C., Moritz, C., Rodrigues, M.T., 2012a. Molecular phylogeny and morphometric analyses reveal deep divergence between Amazonia and Atlantic forest species of Dendrophryniscus. Mol. Phylogenet. Evol. 62, 826–838. Fouquet, A., Santana Cassini, C., Fernando Baptista Haddad, C., Pech, N., Rodrigues, M.T., 2014. Species delimitation, patterns of diversification and historical biogeography of the Neotropical frog genus Adenomera (Anura, Leptodactylidae). J. Biogeog. 41, 855-870. Freitas, M.A., Machado, D.C., Venâncio, N.M., França, D.P.F., Veríssimo, D., 2013. First record for Brazil of the odd anole lizard, Anolis dissimilis Williams, 1965 (Squamata: Polychrotidae) with notes on coloration. Herpetol. Notes 6, 383–385. Frichot, E., Schoville, S.D., Bouchard, G., François, O., 2013. Testing for associations between loci and environmental gradients using latent factor mixed models. Mol. Biol. Evol., 30, 1687-1699. 132

Frichot, E., Mathieu, F., Trouillon, T., Bouchard, G., François, O., 2014. Fast and efficient estimation of individual ancestry coefficients. Genetics 196, 973-983. Frichot, E., François, O., 2015. LEA: an R package for landscape and ecological association studies. Methods Ecol. Evol. 6, 925-929. Gartner, G.E., Gamble, T., Jaffe, A.L., Harrison, A., Losos, J.B., 2013. Left–right dewlap asymmetry and phylogeography of Anolis lineatus on Aruba and Curaçao. Biol. J. Linn. Soc. 110, 409–426. Gehara, M., Crawford, A.J., Orrico, V.G., Rodríguez, A., Lötters, S., Fouquet, A., Barrientos, L.S., Brusquetti, F., De la Riva, I., Ernst, R., Urrutia, G.G., 2014. High levels of diversity uncovered in a widespread nominal taxon: continental phylogeography of the neotropical tree frog Dendropsophus minutus. Plos One 9, e103958. Gernhard, T., 2008. The conditioned reconstructed process. J. Theor. Biol. 253, 769-778. Geurgas, S.R., Rodrigues, M.T., 2010. The hidden diversity of Coleodactylus amazonicus (Sphaerodactylinae, Gekkota) revealed by molecular data. Mol. Phyl. Evol. 54, 583-593. Ghalambor, C.K., Huey, R.B., Martin, P.R., Tewksbury, J.J., Wang, G., 2006. Are mountain passes higher in the tropics? Janzen's hypothesis revisited. Integ. Comp. Biol. 46, 5-17 Glor, R.E., Vitt, L.J., Larson, A., 2001. A molecular phylogenetic analysis of diversification in Amazonian Anolis lizards. Mol. Ecol. 10, 2661-2668. Gould, S.J., Vrba, E.S., 1982. Exaptation—a missing term in the science of form. Paleobiol. 8, 4- 15. Graham, C.H., Ron, S.R., Santos, J.C., Schneider, C.J., Moritz, C., 2004. Integrating phylogenetics and environmental niche models to explore speciation mechanisms in Dendrobatid frogs. Evolution 58(8), 1781-1793. Gray, J.E., 1831. A synopsis of the species of class Reptilia, in: Griffith, E., Pidgeon, E. (Eds.), The animal kingdom arranged in conformity with its organisation by the Baron Cuvier with additional descriptions of all the species hither named, and of many before noticed, vol. 9. Whittaker, Treacher and Co., London. Gronau, I., Hubisz, M.J., Gulko, B., Danko, C.G., Siepel, A., 2011. Bayesian inference of ancient human demography from individual genome sequences. Nature Genetics, 43, 1031-1034. Guindon, S., Gascuel, O., 2003. A simple, fast and accurate method to estimate large phylogenies by maximum-likelihood. Syst. Biol. 52, 696–704. Guo, B., Lu, D., Liao, W.B., Merilä, J., 2016. Genomewide scan for adaptive differentiation along altitudinal gradient in the Andrew's toad Bufo andrewsi. Mol. Ecol., 25, 3884-3900. Haffer, J., 1969. Speciation in Amazonian forest birds. Science 165, 131-137. Harwood, R.H., 1979. The effect of temperature on the digestive efficiency of three species of lizards, Cnemidophorus tigris, Rrhonotus multicarinatus and Sceloporus occidentalis. Comp. Biochem. Physiol. A, 63, 417-433. Hasegawa, M., Kishino, H., Yano, T.A., 1985. Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J. Mol. Evol. 22, 160-174. Heled, J., Drummond, A.J., 2010. Bayesian inference of species trees from multilocus data. Mol. Biol. Evol. 27, 570-580. Hertz, P.E., Huey, R.B., Nevo, E., 1983. Homage to Santa Anita: thermal sensitivity of sprint speed in agamid lizards. Evolution, 37, 1075-1084. Hickerson, M.J., Stahl, E., Takebayashi, N., 2007. msBayes: pipeline for testing comparative phylogeographic histories using hierarchical approximate Bayesian computation. BMC Bioinformatics 8, 268. 133

Hijmans, R.J., Cameron, S.E., Parra, J.L., Jones, P.G., Jarvis, A., 2005. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25, 1965-1978. Ho, S.Y., Phillips, M.J., 2009. Accounting for calibration uncertainty in phylogenetic estimation of evolutionary divergence times. Syst. Biol. 58, 367-380. Hohenlohe, P.A., Bassham, S., Etter, P.D., Stiffler, N., Johnson, E.A., Cresko, W.A., 2010. Population genomics of parallel adaptation in threespine stickleback using sequenced RAD tags. PLoS Genetics, 6, e1000862. Holman, J.A., 1972. Herpetofauna of Calf Creek local fauna (lower Oligocene, Cypress Hills formation) of Saskatchewan. Can. J. Earth Sci. 9, 1612-1631. Holman, J.A., 1987. Some amphibians and reptiles from the Oligocene of northeastern Colorado. Dakoterra 3, 16-21. Hoorn, C., Wesselingh, P., ter Steege, H., Bermudez, M.A., Mora, A., Sevink, J., Sanmartin, I., Sanchez-Meseguer, A., Anderson, C.L., Figueredo, J.P., Jaramillo, C., Riff, D., Negri, F.R., Hooghiemstra, H., Lundberg, J., Stadler, T., Sarkinen, T. Antonelli, A., 2010. Amazonia through time: Andean uplift, climate change, landscape evolution, and biodiversity. Science 330, 927–931. Huang, W., Takebayashi, N., Qi, Y., Hickerson, M.J., 2011. MTML-msBayes: approximate Bayesian comparative phylogeographic inference from multiple taxa and multiple loci with rate heterogeneity. BMC Bioinformatics 12, 1. Huey, R.B., Gilchrist, G.W., Carlson, M.L., Berrigan, D., Serra, L., 2000. Rapid evolution of a geographic cline in size in an introduced fly. Science, 287, 308-309. Hughes, C., Eastwood, R., 2006. Island radiation on a continental scale: exceptional rates of plant diversification after uplift of the Andes. Proc. Nat. Acad. Sci. 103, 10334-10339. Icochea, J., Quispitupac, E., Portilla, A., Ponce, E., 2001. Assessment of amphibians and reptiles of the Lower Urubamba Region, Peru, in: Alonso, A., Dallmeier, F., Campbell, P. (Eds.), Urubamba: The Biodiversity of a Peruvian Rain Forest. Smithsonian Institution, Washington. Jenssen, T.A., Gladson, N.L., 1984. A comparative display analysis of the Anolis brevirostris complex in Haiti. J. Herpetol. 18, 217-230. Jezkova, T., Leal, M., Rodriguez-Robles, J.A., 2009. Living together but remaining apart: comparative phylogeography of Anolis poncensis and A. cooki, two lizards endemic to the aridlands of . Biol. J. Linn. Soc. 96, 617–634. John-Alder, H.B., Barnhart, M.C., Bennett, A.F., 1989. Thermal sensitivity of swimming performance and muscle contraction in northern and southern populations of tree frogs (Hyla crucifer). J. Exp. Biol., 142, 357-372. Johnson, J.A., Watson, R.T., Mindell, D.P., 2005. Prioritizing species conservation: does the Cape Verde kite exist?. Proc. R. Soc. Lond. B 272, 1365-1371. Kawashita-Ribeiro, R.A., Avila, R.W., 2008. Reptilia, Squamata, Polychrus spp.: New record, range extensions, and distribution map in the state of Mato Grosso, Brazil. Check List 4, 362- 365. Kawecki, T.J., Ebert, D., 2004. Conceptual issues in local adaptation. Ecol. Letters, 7, 1225- 1241. Knowles, L.L., Carstens, B.C., 2007. Delimiting species without monophyletic gene trees. Syst. Biol. 56, 887-895. Koehler, G., 2014. Characters of external morphology used in Anolis taxonomy–Definition of terms, advice on usage, and illustrated examples. Zootaxa 3774, 201-257. Kolbe, J.J., Van Middlesworth, P.S., Losin, N., Dappen, N., Losos, J.B., 2012. Climatic niche 134

shift predicts thermal trait response in one but not both introductions of the Puerto Rican lizard Anolis cristatellus to Miami, Florida, USA. Ecol. Evol., 2, 1503-1516. Kozak, K.H., Wiens, J., 2006. Does niche conservatism promote speciation? A case study in North American salamanders. Evolution 60, 2604-2621. Kubatko, L.S., Degnan, J.H., 2007. Inconsistency of phylogenetic estimates from concatenated data under coalescence. Syst. Biol. 56, 17-24. Lambert, S.M., Geneva, A.J., Luke Mahler, D. Glor, R.E., 2013. Using genomic data to revisit an early example of reproductive character displacement in Haitian Anolis lizards. Mol. Ecol. 22, 3981–3995. Lanfear, R., Calcott, B., Ho, S.Y., Guindon, S., 2012. PartitionFinder: combined selection of partitioning schemes and substitution models for phylogenetic analyses. Mol. Biol. Evol. 29, 1695–1701. Lazell Jr., J.D., 1965. An Anolis (Sauria, Iguanidae) in amber. J. Paleontol. 39, 379-382. Lazell Jr., J.D., 1969. The genus Phenacosaurus (Sauria: Iguanidae). Breviora 325, 1–24. Leal, M., Gunderson, A.R., 2012. Rapid change in the thermal tolerance of a tropical lizard. Am. Nat. 180, 815-822. Lenormand, T., 2002. Gene flow and the limits to natural selection. Trends Ecol. Evol., 17, 183- 189. Linnaeus, C., 1758. Systema naturae per regna tria naturae, secundum classes, ordines, genera, species, cum characteribus, differentiis, synonymis, locis. Tomus I. Editio decima, reformata. Laurentii Salvii, Stockholm. Losos, J.B., 1985. An experimental demonstration of the species-recognition role of Anolis dewlap color. Copeia 1985, 905-910. Losos, J.B., 2009. Lizards in an Evolutionary Tree: Ecology and Adaptive Radiation of Anoles. University of California Press, Berkeley. Losos, J.B., Woolley, M.L., Mahler, D.L., Torres-Carvajal, O., Crandell, K.E., Schaad, E.W., Narváez, A.E., Ayala-Varela, F., Herrel, A., 2012. Notes on the natural history of the little- known Ecuadorian horned anole, Anolis proboscis. Breviora 531, 1-17. Lukoschek, V., Keogh, J.S., Avise, J.C., 2012. Evaluating fossil calibrations for dating phylogenies in light of rates of molecular evolution: a comparison of three approaches. Syst. Biol. 61, 22-43. Lundberg, J.G., Marshall, L.G., Guerrero, J., Horton, B., Malabarba, M.C.S.L., Wesselingh, F.P., 1998. The stage for Neotropical fish diversification: a history of tropical South American rivers, in: Malabarba, L.R., Reis, R.E., Yari, R.P., Lucena, Z.M., Lucena, C.A.S. (Eds.), Phylogeny and Classification of Neotropical Fishes. Edipucrs, Porto Alegre. Macedonia, J.M., Evans, C.S., Losos, J.B., 1994. Male Anolis lizards discriminate video- recorded conspecific and heterospecific displays. Anim. Behav. 47, 1220-1223. Macedonia, J.M., Stamps, J.A., 1994. Species recognition in Anolis grahami (Sauria, Iguanidae): evidence from responses to video playbacks of conspecific and heterospecific displays. Ethology 98, 246-264. McKay, B.D., Zink, R.M., 2010. The causes of mitochondrial DNA gene tree paraphyly in birds. Mol. Phyl. Evol. 54, 647-650. Melo-Sampaio, P.R., Melo, B.L.A.S., Silva, A., Maciel, J.M.L., Nogueira, M., Arruda, S.M., Lima, L., Silva, J.C., Matos, L., 2013. Geographic distribution: Anolis dissimilis. Herpetol. Rev. 44, 473. Mulcahy, D.G., Noonan, B.P., Moss, T., Townsend, T.M., Reeder, T.W., Sites Jr, J.W., Wiens, 135

J.J., 2012. Estimating divergence dates and evaluating dating methods using phylogenomic and mitochondrial data in squamate reptiles. Mol. Phyl. Evol. 65, 974-991. Muñoz, M.M., Stimola, M.A., Algar, A.C., Conover, A., Rodriguez, A.J., Landestoy, M.A., Bakken, G.S. and Losos, J.B., 2014. Evolutionary stasis and lability in thermal physiology in a group of tropical lizards. Proc. R. Soc. B 281, 20132433. Muñoz, M.M., Crawford, N.G., McGreevy, T.J., Messana, N.J., Tarvin, R.D., Revell, L.J., Zandvliet, R.M., Hopwood, J.M., Mock, E., Schneider, A.L., Schneider, C.J., 2013. Divergence in coloration and ecological speciation in the Anolis marmoratus species complex. Mol. Ecol. 22, 2668–2682. Myers, C.W., Donnelly, M.A., 1996. A new herpetofauna from Cerro Yaví, Venezuela: first results of the Robert G. Goelet American Museum-TERRAMAR expedition to the northwestern Tepuis. Am. Mus. Novit. 3172, 1–56. Myers, C.W., Williams, E.E., McDiarmid, R.W., 1993. A new anoline lizard (Phenacosaurus) from the highland of Cerro de la Neblina, southern Venezuela. Am. Mus. Novit. 3070, 1–15. Myers, G.S., Carvalho, A.L., 1945. A strange leaf-nosed lizard of the genus Anolis from Amazonia. Bol. Mus. Nac. Sér. Zool. 43, 1-22. Navas, C.A., 2002. Herpetological diversity along Andean elevational gradients: links with physiological ecology and evolutionary physiology. Comp. Biochem. Physiol. A 133, 469- 485. Ng, J., Glor, R.E., 2011. Genetic differentiation among populations of a Hispaniolan trunk anole that exhibit geographical variation in dewlap color. Mol. Ecol. 20, 4302–4317. Nicholson, K.E., 2002. Phylogenetic analysis and a test of the current infrageneric classification of Norops (Beta Anolis). Herpetol. Monogr. 16, 93–120. Nicholson, K.E., Crother, B.I., Guyer, C., Savage, J.M., 2012. It is time for a new classification of anoles (Squamata: Dactyloidae). Zootaxa 3477, 1-108. Nicholson, K.E., Glor, R.E., Kolbe, J.J., Larson, A., Hedges, S.B., Losos, J.B., 2005. Mainland colonization by island lizards. J. Biogeogr. 32, 929–938. Nosil, P., 2012. Ecological speciation. Oxford University Press. Oaks, J.R., 2014. An improved approximate-Bayesian model-choice method for estimating shared evolutionary history. BMC Evol. Biol. 14, 1. Oaks, J.R., Sukumaran, J., Esselstyn, J.A., Linkem, C.W., Siler, C.D., Holder, M.T., Brown, R.M., 2013. Evidence for climate-driven diversification? a caution for interpreting abc inferences of simultaneous historical events. Evolution, 67, 991-1010. Omland, K.E., Tarr, C.L., Boarman, W.I., Marzluff, J.M., Fleischer, R.C., 2000. Cryptic genetic variation and paraphyly in ravens. Proc. R. Soc. Lond. B 267, 2475-2482. Padial, J.M., Grant, T., Frost, D.R., 2014. Molecular systematics of terraranas (Anura: Brachycephaloidea) with an assessment of the effects of alignment and optimality criteria. Zootaxa 3825, 1-132. Paradis, E., Claude J., Strimmer K., 2004. APE: analyses of phylogenetics and evolution in R language. Bioinformatics 20, 289–290. Parham, J.F., Donoghue, P.C., Bell, C.J., Calway, T.D., Head, J.J., Holroyd, P.A., Inoue, J.G., Irmis, R.B., Joyce, W.G., Ksepka, D.T., Patané, J.S.L., Smith, N.D., Tarver, J.E., van Tuinen, M., Yang, Z., Angielczyk, K.D., Greenwood, J.M., Hipsley, C.A., Jacobs, L., Makovicky, P.J., Müller, J., Smith, K.T., Theodor, J.M., Warnock, R.C.M., 2011. Best practices for justifying fossil calibrations. Syst. Biol. 61, 346-359. Paz, A., Ibáñez, R., Lips, K.R., Crawford, A.J., 2015. Testing the role of ecology and life history 136

in structuring genetic variation across a landscape: a trait-based phylogeographic approach. Mol. Ecol. 24, 3723-3737. Pellegrino, K.C., Rodrigues, M.T., Harris, D.J., Yonenaga-Yassuda, Y., Sites, J.W., 2011. Molecular phylogeny, biogeography and insights into the origin of parthenogenesis in the Neotropical genus Leposoma (Squamata: Gymnophthalmidae): Ancient links between the Atlantic Forest and Amazonia. Mol. Phyl. Evol. 61, 446-459. Percequillo, A.R., Weksler, M., Costa, L.P., 2011. A new genus and species of rodent from the Brazilian Atlantic Forest (Rodentia: Cricetidae: Sigmodontinae: Oryzomyini), with comments on oryzomyine biogeography. Zool. J. Linn. Soc. 161, 357-390. Peters, J.A., Orces, V.G., 1956. A third leaf-nosed species of the lizard genus Anolis from South America. Breviora 62, 1-8. Peters, W.C.H., 1863. Über einige neue Arten der Saurier-Gattung Anolis. Monatsber. Königl. Akad. Wiss. Berlin. 1863, 135-149. Petren, K., Grant, P.R., Grant, B.R., Keller, L.F., 2005. Comparative landscape genetics and the adaptive radiation of Darwin's finches: the role of peripheral isolation. Mol. Ecol. 14, 2943- 2957. Pimm, S.L., Jenkins, C.N., Joppa, L.N., Roberts, D.L., Russell, G.J., 2010. How many endangered species remain to be discovered in Brazil. Nat. Conserv. 8, 71-77. Pirani, R.M., Mângia, S., Santana, D.J., Assis, B.D., Feio, R.N., 2010. Rediscovery, distribution extension and natural history notes of Hylodes babax (Anura, Hylodidae) with comments on Southeastern Brazil Biogeography. S. Am. J. Herpetol. 5, 83-88. Poe, S., 1998. Skull characters and the cladistic relationships of the Hispaniolan dwarf twig Anolis. Herpetol. Monog. 12, 192-236. Poe, S., 2004. Phylogeny of anoles. Herpetol. Monogr. 18, 37-89. Poe, S., 2013. 1986 Redux: New genera of anoles (Squamata: Dactyloidae) are unwarranted. Zootaxa 3626, 295-299. Poe, S., Ayala, F., Latella, I.M., Kennedy, T.L., Christensen, J.A., Gray, L.N., Blea, N.J., Armijo, B.M., Schaad, E.W., 2012. Morphology, phylogeny, and behavior of Anolis proboscis. Breviora 530, 1-11. Poe, S., Latella, I., Ayala-Varela, F., Yañez-Miranda, C., Torres-Carvajal, O., 2015. A new Species of Phenacosaur Anolis (Squamata; Iguanidae) from Peru and a comprehensive phylogeny of Dactyloa-clade Anolis based on new DNA sequences and morphology. Copeia 103, 639-650. Poe, S., Yañez-Miranda, C., 2007. A new species of phenacosaur Anolis from Peru. Herpetologica 63, 219–223. Pons, J., Barraclough, T.G., Gomez-Zurita, J., Cardoso, A., Duran, D.P., Hazell, S., Kamoun, S., Sumlin, W.D., Vogler, A.P., 2006. Sequence-based species delimitation for the DNA taxonomy of undescribed insects. Syst. Biol. 55, 595-609. Portik, D.M., Wood, P.L., Grismer, J.L., Stanley, E.L., Jackman, T.R., 2012. Identification of 104 rapidly-evolving nuclear protein-coding markers for amplification across scaled reptiles using genomic resources. Conserv. Gen. Res. 4, 1-10. Prates, I., Angilleta Jr, M.J., Wilson, R.S., Niehaus, A.C., Navas, C.A., 2013. Dehydration hardly slows hopping toads (Rhinella granulosa) from xeric and mesic environments. Physiol. Biochem. Zool. 86, 451-457. Prates, I., Melo-Sampaio, P.R., de Oliveira Drummond, L., Teixeira, M., Rodrigues, M.T. and Carnaval, A.C., 2017. Biogeographic links between southern Atlantic Forest and western 137

South America: Rediscovery, re-description, and phylogenetic relationships of two rare montane anole lizards from Brazil. Mol. Phyl. Evol. 113, 49-58. Prates, I., Rodrigues, M.T., Melo-Sampaio, P.R., Carnaval, A.C., 2015. Phylogenetic relationships of Amazonian anole lizards (Dactyloa): taxonomic implications, new insights about phenotypic evolution and the timing of diversification. Mol. Phylogenet. Evol. 82, 258–268. Prates, I., Xue, A.T., Brown, J.L., Alvarado-Serrano, D.F., Rodrigues, M.T., Hickerson, M.J., Carnaval, A.C., 2016a. Inferring responses to climate dynamics from historical demography in neotropical forest lizards. Proc. Nat. Acad. Sci. 113, 7978-7985. Prates, I., Rivera, D., Rodrigues, M.T., Carnaval, A.C., 2016b. A mid-Pleistocene rainforest corridor enabled synchronous invasions of the Atlantic Forest by Amazonian anole lizards. Mol. Ecol. 25, 5174-5186. R Core Team, 2017. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna. Retrieved from http://www.R-project.org/. Reid, N.M., Carstens, B.C., 2012. Phylogenetic estimation error can decrease the accuracy of species delimitation: a Bayesian implementation of the general mixed Yule-coalescent model. BMC Evol. Biol. 12, 196. Rellstab, C., Gugerli, F., Eckert, A.J., Hancock, A.M., Holderegger, R., 2015. A practical guide to environmental association analysis in landscape genomics. Mol. Ecol. 24, 4348-4370. Ribeiro-Junior, M.A., 2015. Catalogue of distribution of lizards (Reptilia: Squamata) from the Brazilian Amazonia. I. Dactyloidae, Hoplocercidae, Iguanidae, Leiosauridae, Polychrotidae, Tropiduridae. Zootaxa 3983, 1-110. Rocha, P.A., Ferrari, S.F., Feijó, A., Gouveia, S.F., 2015. Zoogeography of South American Forest-Dwelling Bats: Disjunct Distributions or Sampling Deficiencies?. PloS One 10, e0133276. Rodrigues, M.T., 1988. A new anole of the punctatus group from Central Amazonia (Sauria, Iguanidae). Pap. Avulsos Zool. 36, 333-336. Rodrigues, M.T., Bertolotto, C.E.V., Amaro, R.C., Yonenaga-Yassuda, Y., Freire, E.M.X., Pellegrino, K.C.M., 2014. Molecular phylogeny, species limits, and biogeography of the Brazilian endemic lizard genus Enyalius (Squamata: Leiosauridae): An example of the historical relationship between Atlantic Forests and Amazonia. Mol. Phyl. Evol. 81, 137-146. Rodrigues, M.T., Cassimiro, J., Pavan, D., Curcio, F.F., Verdade, V.K., Pellegrino, K.C.M., 2009. A new genus of microteiid lizard from the Caparaó mountains, southeastern Brazil, with a discussion of relationships among Gymnophthalminae (Squamata). Am. Mus. Nov. 3673, 1-27. Rodrigues, M.T., Pellegrino, K.C.M., Dixo, M., Verdade, V.K., Pavan, D., Argolo, A.J.S., Sites Jr, J.W., 2007. A new genus of microteiid lizard from the Atlantic Forests of state of Bahia, Brazil, with a new generic name for Colobosaura mentalis, and a discussion of relationships among the Heterodactylini (Squamata, Gymnophthalmidae). Am. Mus. Nov. 3565, 1-27. Rodrigues, M.T., Teixeira Jr., M., Recoder, R.S., Dal Vechio, F., Damasceno, R., Pellegrino, K.C.M., 2013. A new species of Leposoma (Squamata: Gymnophthalmidae) with four fingers from the Atlantic Forest central corridor in Bahia, Brazil. Zootaxa 3635, 459–475. Rodrigues, M.T., Xavier, V., Skuk, G., Pavan, D., 2002. New specimens of Anolis phyllorhinus (Squamata, Polychrotidae): the first female of the species and of proboscid anoles. Pap. Avulsos Zool. 42, 363-380. Ronquist, F., Teslenko, M., van der Mark, P., Ayres, D.L., Darling, A., Höhna, S., Larget, B., 138

Liu, L., Suchard, M.A., Huelsenbeck, J.P., 2012. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 61, 539-542. Santos, A.M., Cavalcanti, D.R., Silva, J.M.C.D., Tabarelli, M., 2007. Biogeographical relationships among tropical forests in north-eastern Brazil. J. Biogeog. 34, 437-446. Santos, J.C., Coloma, L.A., Summers, K., Caldwell, J.P., Ree, R., Cannatella, D.C., 2009. Amazonian amphibian diversity is primarily derived from late Miocene Andean lineages. PLoS Biol. 7, e1000056. Scheffers, B.R., Joppa, L.N., Pimm, S.L., Laurance, W.F., 2012. What we know and don’t know about Earth's missing biodiversity. Trends Ecol. Evol. 27, 501-510. Sexton, J.P., Hangartner, S.B., Hoffmann, A.A., 2014. Genetic isolation by environment or distance: which pattern of gene flow is most common?. Evolution, 68, 1-15. Sigmund, W.R., 1983. Female preference for Anolis carolinensis males as a function of dewlap color and background coloration. J. Herpetol. 17, 137-143. Smith, K.T., 2006. A diverse new assemblage of late Eocene squamates (Reptilia) from the Chadron formation of North Dakota, USA. Palaeontol. Electron. 9, 1-44. Smith, K.T., 2009. A new lizard assemblage from the earliest Eocene (zone Wa0) of the Bighorn Basin, Wyoming, USA: biogeography during the warmest interval of the Cenozoic. J. Syst. Palaeontol. 7, 299-358. Smith, K.T., 2011. The long-term history of dispersal among lizards in the early Eocene: new evidence from a microvertebrate assemblage in the Bighorn basin of Wyoming, USA. Palaeontology 54, 1243-1270. Sobral-Souza, T., Lima-Ribeiro, M.S., Solferini, V.N., 2015. Biogeography of Neotropical Rainforests: past connections between Amazon and Atlantic Forest detected by ecological niche modeling. Evol. Ecol. 29, 643-655. Stapley, J., Wordley, C., Slate, J., 2011. No evidence of genetic differentiation between anoles with different dewlap color patterns. J. Hered. 102, 118-124. Stephens, M., Donnelly, P., 2003. A comparison of Bayesian methods for haplotype reconstruction from population genotype data. Am. J. Hum. Genet. 73, 1162-1169. Suchard, M.A., Weiss, R.E., Sinsheimer, J.S., 2001. Bayesian selection of continuous-time Markov chain evolutionary models. Mol. Biol. Evol. 18, 1001-1013. Sukumaran, J., Holder, M.T., 2010. DendroPy: A Python library for phylogenetic computing. Bioinformatics 26, 1569-1571. Sullivan, J., Joyce, P., 2005. Model selection in phylogenetics. Annu. Rev. Ecol. Evol. Syst. 36, 445-466. Sullivan, R.M., Holman., J.A., 1996. Squamata, in: Prothero, D.R., Emry, R.J. (Eds.), The terrestrial Eocene-Oligocene transition in North America. Cambridge University Press, Cambridge, pp. 354-372. Tajima, F., 1989. Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics 123, 585-595. Tamura, K., Nei, M., 1993. Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Mol. Biol. Evol. 10, 512–526. Teixeira Jr., M., Dal Vechio, F., Recoder, R.S., Carnaval, A.C., Strangas, M., Damasceno, R.P., Sena, M.D., Rodrigues, M.T., 2012. Two new species of marsupial tree-frogs genus Gastrotheca Fitzinger, 1843 (Anura, Hemiphractidae) from the Brazilian Atlantic Forest. Zootaxa 3437, 1-23. Teixeira Jr., M., Recoder, R.S., Amaro, R.C., Damasceno, R.P., Cassimiro, J., Rodrigues, M.T., 139

2013. A new Crossodactylodes Cochran, 1938 (Anura: Leptodactylidae: Paratelmatobiinae) from the highlands of the Atlantic Forests of southern Bahia, Brazil. Zootaxa 3702, 459-472. Thorpe, R.S., Stenson, A.G., 2003. Phylogeny, paraphyly and ecological adaptation of the color and pattern in the Anolis roquet complex on Martinique. Mol. Ecol. 12, 117-132. Tollis, M., Ausubel, G., Ghimire, D., Boissinot, S., 2012. Multi-locus phylogeographic and population genetic analysis of Anolis carolinensis: historical demography of a genomic model species. PLoS One 7, e38474. Tonini, J.F.R., Mendonça, I.S., Coutinho, A.B., Gasparini, J.L., 2011. Anurans from Costa Bela, state of Espírito Santo, south-eastern Brazil: inventory at an urban area and the rediscovery of Allobates in the state. Herpetol. Notes 4, 435-444. Townsend, T.M., Alegre, R.E., Kelley, S.T., Wiens, J.J., Reeder, T.W., 2008. Rapid development of multiple nuclear loci for phylogenetic analysis using genomic resources: an example from squamate reptiles. Mol. Phyl. Evol. 47, 129-142. Townsend, T.M., Mulcahy, D.G., Noonan, B.P., Sites Jr, J.W., Kuczynski, C.A., Wiens, J.J., Reeder, T.W., 2011. Phylogeny of iguanian lizards inferred from 29 nuclear loci, and a comparison of concatenated and species-tree approaches for an ancient, rapid radiation. Mol. Phyl. Evol. 61, 363-380. Ugueto, G.N., Rivas, G., Barros, T., Sánchez-Pacheco, S.J., García-Perez, J., 2007. A revision of the Venezuelan Anoles I: A new Anolis species from the Andes of Venezuela with the redescription of Anolis jacare Boulenger 1903 (Reptilia: Polychrotidae) and the clarification of the status of Anolis nigropunctatus Williams 1974. Zootaxa 1501, 1–30. Vaidya, G., Lohman, D.J., Meier, R., 2011. SequenceMatrix: concatenation software for the fast assembly of multi-gene datasets with character set and codon information. Cladistics 27, 171–180. Van Damme, R., Bauwens, D., Castilla, A.M., Verheyen, R.F., 1989. Altitudinal variation of the thermal biology and running performance in the lizard Podarcis tiliguerta. Oecologia 80, 516-524. Vanzolini, P.E., Williams, E.E., 1970. South American anoles: the geographic differentiation and evolution of the Anolis chrysolepis species group (Sauria, Iguanidae). Arq. Zool. 19, 1-176. Vitt, L.J., Cristina, T., Avila-Pires, S., Zani, P.A., Espósito, M.C., 2002. Life in shade: the ecology of Anolis trachyderma (Squamata: Polychrotidae) in Amazonian Ecuador and Brazil, with comparisons to ecologically similar anoles. Copeia, 2002, 275-286. Vitt, L.J., Avila-Pires, T.C.S., Espósito, M.C., Sartorius, S.S., Zani, P.A., 2003. Sharing Amazonian rain-forest trees: Ecology of Anolis punctatus and Anolis transversalis (Squamata: Polychrotidae). J. Herpetol. 276-285. Weir, B.S., Cockerham, C.C., 1984. Estimating F-statistics for the analysis of population structure. Evolution 1358-1370. Werner, F., 1894. Über einige Novitäten der herpetologischen Sammlung des Wiener zoolog. vergl. anatom. Instituts. Zool. Anz. 17, 155-157. Wesselingh, F.P., Salo, J.A., 2006. A Miocene perspective on the evolution of the Amazonian biota. Scripta Geol. 133, 439-458. Wiens, J.J., Hutter, C.R., Mulcahy, D.G., Noonan, B.P., Townsend, T.M., Sites, J.W., Reeder, T.W., 2012. Resolving the phylogeny of lizards and snakes (Squamata) with extensive sampling of genes and species. Biol. Lett. 8, 1043-1046. Williams, E.E., 1965. South American Anolis (Sauria: Iguanidae): two new species of the punctatus group. Breviora 233, 1–15. 140

Williams, E.E., 1974. South American Anolis: Three new species related to Anolis nigrolineatus and A. dissimilis. Breviora 422, 1-15. Williams, E.E., 1976. South American anoles: the species groups. Pap. Avulsos Zool. 29, 259– 268. Williams, E.E., 1979. South American anoles: the species groups. 2. The proboscis anoles (Anolis laevis group). Breviora 449, 1-19. Williams, E.E., 1982. Three new species of the Anolis punctatus complex from Amazonian and inter-Andean Colombia, with comments on the eastern members of the punctatus group. Breviora 467, 1-38. Williams, E.E., 1992. New or problematic Anolis from Colombia: Anolis lamari, a new anole from the Cordillera Oriental of Colombia, with a discussion of tigrinus and punctatus species group boundaries. Breviora 495, 1-24. Williams, E.E., Praderio, M.J., Gorzula, S., 1996. A phenacosaur from Chimantá Tepui, Venezuela. Breviora 506, 1–15. Williams, E.E., Vanzolini, P.E., 1980. Notes and biogeographic comments on anoles of Brazil. Pap. Avulsos Zool. 34, 99-108. Xue, Y., Zhang, H., Wang, H., Hu, J., Du, M., Zhu, M.J., 2014. Host inflammatory response inhibits Escherichia coli O157: H7 adhesion to gut epithelium through augmentation of mucin expression. Infect. Immun., 82, 1921-1930. Yatkola, D.A., 1976. Mid-Miocene lizards from western Nebraska. Copeia 1976, 645-654. Yánez-Muñoz, M.H., Urgilés, M.A., Altamirano, M., Cáceres, S.R., 2010. Redescripción de Anolis proboscis Peters & Orcés (Reptilia: Polychrotidae), con el descubrimiento de las hembras de la especie y comentarios sobre su distribución y taxonomía. Avances Cienc. Ing. 2, B7–B15. Zaher, H., Aguiar, E., Pombal Jr., J.P., 2005. gaigeae (Cochran, 1938) re- discovered (Amphibia, Anura, Leptodactylidae). Arq. Mus. Nac. 63, 321-328. Zamudio, K.R., Bell, R.C., Mason, N.A., 2016. Phenotypes in phylogeography: Species’ traits, environmental variation, and vertebrate diversification. Proc. Nat. Acad. Sci. 113, 8041- 8048. Zheng, Y., Peng, R., Kuro-o, M., Zeng, X., 2011. Exploring patterns and extent of bias in estimating divergence time from mitochondrial DNA sequence data in a particular lineage: a case study of salamanders (Order Caudata). Mol. Biol. Evol. 28, 2521-2535. Zwickl, D.J., 2006. Genetic algorithm approaches for the phylogenetic analysis of large biological sequence datasets under the maximum likelihood criterion. Ph.D. dissertation, The University of Texas, Austin.

141