<<

University of Mississippi eGrove

Electronic Theses and Dissertations Graduate School

2019

Looking into the Past: The Diversification of in the Neotropical Savannas

Ísis da Costa Arantes University of Mississippi

Follow this and additional works at: https://egrove.olemiss.edu/etd

Part of the Biology Commons

Recommended Citation da Costa Arantes, Ísis, "Looking into the Past: The Diversification of Amphibians in the Neotropical Savannas" (2019). Electronic Theses and Dissertations. 1567. https://egrove.olemiss.edu/etd/1567

This Dissertation is brought to you for free and open access by the Graduate School at eGrove. It has been accepted for inclusion in Electronic Theses and Dissertations by an authorized administrator of eGrove. For more information, please contact [email protected]. LOOKING INTO THE PAST: THE DIVERSIFICATION OF AMPHIBIANS IN THE NEOTROPICAL SAVANNAS

A dissertation

submitted to the graduate faculty

in partial fulfillment of the requirements for the

Degree of Doctor of Philosophy

in the Biology Department

The University of Mississippi

By

ÍSIS DA COSTA ARANTES, B.S.; M.S.

May, 2019

Copyright Ísis da Costa Arantes 2019 ALL RIGHTS RESERVED

ii

ABSTRACT

Aim: Reveal the evolutionary processes that contributed to biotic diversification in the Cerrado savanna using amphibians as a model. Evolutionary patterns were investigated by comparing the phylogeographic and biogeographic history of biome-specific yet widely distributed (Chiasmocleis albopunctata, rubicundulus and Physalaemus nattereri).

Location: Cerrado region, central .

Methods: I sampled thousands of loci randomly distributed throughout the genomes of all three species. I applied phylogenetic, phylogeographic, demographic, and coalescent species delimitation methods to these molecular data, in combination with species distribution modeling for the past, in order to resolve questions of evolutionary history, taxonomic diversity, species boundaries, and test hypotheses that implicate climatic (stable/unstable) and geomorphological events (plateau/valley) in the formation of this diversity.

Results: The C. albopuntata species group is distributed in the Bolivian and Brazilian savannas and Chaco and it is comprised of at least tree species C. albopunctata, C. mehelyi, and a composite lineage that includes: C. bicegoi, C. centralis, and C. sp. The Cerrado species C. centralis, D. rubicundulus and P. nattereri were each found to have three distinct, genetic populations widely distributed in the Neotropical savannas. No patterns of genetic differentiation related to geomorphology (plateau vs. valley populations) nor climatic stability in any of the species was observed. The demographic models indicate the presence of migration between populations, demonstrating that there are no important geographic barriers to gene flow.

Conclusions: The results highlight the likely influence of the Atlantic as the source for C. albopunctata species group inhabiting the dry open areas of the Neotropical savannas. The further uplift of the Brazilian shield during the Pliocene and Pleistocene was an important event for diversification of C. centralis, D. rubicundulus and P. nattereri, and the isolation by distance also promoted the differentiation among the populations of C. centralis, D. rubicundulus and P. nattereri, in combination with the climatic fluctuations during the Quaternary.

iii

Key-words: Cerrado, anurans, comparative phylogeography, biogeography, species delimitation, open areas, , , Leptodactylidae.

iv

DEDICATED TO

My parents, Izabel and Sócrates, for all the parental care and love.

My son, Benjamin, and my love, Kevin, for all the support and love during my PhD.

v

ACKNOWLEDGMENTS

I would like to thank my family for their support when I decided to pursue a Ph.D. abroad. The decision to leave my family, my friends, my country and culture was one of the hardest things I have done in my life. I would like to especially thank my parents, Izabel and

Sócrates (in memoriam), and my siblings, Taís, Sócrates, Rebeca, and Henrique. I would like to thank my son, Benjamin, for making me stronger than I was, and my fiancé, Kevin, for all the love and support during the final stretch of my Ph.D. Also, I’m very grateful to be part of the

Felker family and I want to thank them for all of their help with Ben.

I want to express my infinite gratitude to my friends, family, and the people that I did not even know that helped me after the car wreck during my first day of fieldwork. I would like to express my eternal gratitude to: Brigadeiro Leônidas, Brigadeiro Rogério Gammerdinger Veras,

Astrid Studart Corrêa, Dr. Carlos Carmona, Marcela Ayub Brasil, Eduardo Xavier Barreto

Junior, Dr. Ângelo Augusto Bongiolo Ganeo, Dr. Guarino Colli, Cecília Vieira, Sr. Hervaldo

Sampaio Carvalho, Sr. Fábio Gondim, Mônica Gondim, Dr. Mário Sallenave, the team from

Brazilian emergency medical service (SAMU), Fabrícius Domingos, Marina Scalon, Laís

Veludo, Iedja Galdino, Bernardo Costa, Lorrainy Bartason, Taís Arantes, Sócrates Arantes,

Ayrton Peres, Paula, Amanda de Sena, Ana Carolina, Bárbara Zimbres, Cintia Carla, Daniel

Velho, Gabriel Horta, Guilherme Santoro, Gustavo Vieira, Isaque Medeiros, Maíra Costa,

Marcelo Luiz, Mariana Caixeta, Poliene Martins, Rafael de Brito, Renata Dias, Reuber Brandão,

Roger Maia, Samuel Enrique, and Taissa Villas Boas. I will always keep all of you in my heart

vi

and thoughts! Also, I would like to thank everyone that thought about me, prayed for my recovery and sent me messages, it helped me to keep positive thoughts and a strong belief that I was going to leave the hospital walking.

I would like to give a special thanks to Dr. Guarino Colli for contributing substantially to the completion of this work. His contributions ensured the successful completion of my field and laboratory work, and also facilitated the logistics of the fieldwork in (before and after my car accident), including collection permits, transportation, and support from his laboratory. In addition, I’m very grateful for having such amazing and supportive advisor, Dr. Brice Noonan. I would like to thank him for all the support during my Ph.D., helping me during my lab work and for being a good friend. Also, I want to thank my committee members for reading and improving my work: Dr. Ryan Garrick, Dr. Jason Hoeksema, and Dr. Louis Zachos. I would like to thank

Benjamin Pharr for all the bioinformatic support and the help with analyses. I would like to thank my collaborators for reading and giving important insights to this work: Mariana

Vasconcellos, Guarino Colli, Rafael de Sá, and João Tonini.

I would like to thank all the collections, curators, and collection managers for sending the tissue samples to make this work possible. I’m very thankful to: Dr. Natan Maciel (Coleção

Zoológica - UFG), Dr. Célio Haddad (collection CFBH), Dr. Adrian Garda (Laboratório de

Anfíbios e Répteis - UFRN), Dr. José Pombal Jr. (Museu Nacional do Rio de Janeiro - UFRJ),

Dr. Luciana Nascimento (Museu de Ciências Naturais - PUC Minas), Dr. Ana Prudente (Museu

Paraense Emílio Goeldi), Dr. Miguel Rodrigues (Coleção de Tecidos de Vertebrados - USP), Dr.

Felipe Curcio (Coleção Zoológica - UFMT), Dr. Hussam Zaher (Museu de Zoologia – USP), Dr.

vii

Francisco Brusquetti (Collección Biológica Arnoldo da Winkelried Bertoni Herpetología - IIBP),

Dr. Nicolás Pelegrin (Instituto de Diversidad y Ecología - CONICET) and Dr. Martin

Jansen (The Senckenberg Research and Nature Museum Frankfurt - SMF) for providing samples from different countries in South America. I’m especially grateful for the Herpetological

Collection of the Universidade de Brasília (CHUNB) for providing me logistic support and allowing me to send the samples to the United States.

I would like to thank the Brazilian Agency Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES # 001181/2013-00) for the financial support during the first four years of my Ph.D., and the Graduate school of the University of Mississippi for the Dissertation

Fellowship. I would like to thank Instituto Chico Mendes de Conservação da Biodiversidade

(ICMBio # 50679-3), Instituto Brasília Ambiental (IBRAM # 001/2016- SUGAP/IBRAM), and

Comissão de Ética de Uso Animal (CEUA - UNB # 66732/2016) for issuing the fieldwork licenses.

I’m grateful for the staff of the Biology department for the all the help, especially Cindy

Rimoldi and Kim Byrd. I would like to thank my lab mates for all of their help with the molecular analysis and for being there for me when I needed: Marcella Santos, Renan Bosque,

Vivian Trevine, Andrew Snyder, Tim Colston, JP Lawrence, Stuart Nielsen, and Megan Smith.

Also, I’m very grateful for my friends that collected with me during my fieldwork: João Pantoja,

Gabriel Caputo, Carolina Azevedo. I would like to thank Cassandra Shields for the help with my lab work.

viii

“Sonho que se sonha só, é só um sonho que se sonha só, mas sonho que se sonha junto é realidade”. Obrigada aos que sonharam comigo!

“A dream that you dream alone, it’s just a dream that you dream alone, but a dream that you dream together becomes reality.” Thank you for dreaming about it with me!

ix

TABLE OF CONTENTS

LIST OF TABLES ...... XII

LIST OF FIGURES ...... XIII

GENERAL INTRODUCTION ...... 1 The Cerrado: Regional setting ...... 2 Biodiversity and conservation status ...... 3 Geomorphologic events and climatic fluctuation ...... 4 Biogeography of Cerrado Amphibians ...... 12 Anurans as a Model of Cerrado Diversification ...... 13 Objectives ...... 15 FIGURES ...... 17

CHAPTER 1 ...... 22

SPECIES DELIMITATION OF CHIASMOCLEIS ALBOPUNTATA SPECIES COMPLEX ...... 22

INTRODUCTION ...... 22 Objective ...... 25 Methods ...... 26 Sampling ...... 26 Molecular methods ...... 28 Phylogenetics, Phylogeography, and Population structure ...... 31 Divergence time ...... 32 Species delimitation ...... 33 Maps ...... 34 Results ...... 34 3RADseq processing ...... 35 Phylogeography ...... 35 Divergence time ...... 36 Species delimitation ...... 36 Discussion ...... 37 Conclusion ...... 41 FIGURES ...... 43

CHAPTER 2 ...... 50

x

COMPARATIVE PHYLOGEOGRAPHY OF THREE BRAZILIAN SAVANNA ...... 50

INTRODUCTION ...... 50 Objective ...... 53 Hypothesis to be tested ...... 54 Predictions ...... 54 Methods ...... 55 Sampling ...... 55 Molecular methods ...... 60 Phylogenetics, Phylogeography, and Population structure ...... 63 Divergence time and Demographic history ...... 64 Population genetic diversity and isolation by distance (IBD) ...... 65 Species distribution models (SDMs) ...... 66 Analysis Molecular of Variance (AMOVA) ...... 68 Maps ...... 69 Results ...... 69 3RADseq processing ...... 69 Phylogenetics, Phylogeography and population structure ...... 70 Divergence time and Demographic history ...... 72 Population genetic diversity and isolation by distance (IBD) ...... 73 Species distribution modeling ...... 74 Analysis Molecular of Variance (AMOVA) ...... 76 Discussion ...... 78 Conclusion ...... 83 FIGURES ...... 85

LIST OF REFERENCES ...... 103

VITA ...... 116

xi

LIST OF TABLES

Table 1: Hypotheses of Cerrado diversification and their predicted effects on the evolutionary history of herpetofauna...... 11 Table 2: Sampling localities of Chiasmocleis albopunctata group specimens used in this study. The number (#) corresponds to the localities on the map (Fig. 10). The Code is used as the locality code in the phylogenetic trees presented in the results (Fig. 6)...... 27 Table 3: Output of iPyRAD for number of total filtered loci and SNPs including and not including outgroups...... 31 Table 4: Results for BFD* species delimitation model selection...... 37 Table 5: Population localities of Chiasmocleis centralis samples used in this study. The number (#) corresponds to the localities in the map (Fig. 21). The Code is used as the locality code and the colors represent the population cluster in the phylogenetic trees presented in the results (Fig. 14)...... 57 Table 6: Population localities of Dendropsophus rubicunculus samples used in this study. The number (#) corresponds to the localities in the map (Fig. 22). The Code is used as the locality code and the colors represent the population cluster in the phylogenetic trees presented in the results (Fig. 15)...... 58 Table 7: Population localities of Physalaemus nattereri samples used in this study. The number (#) corresponds to the localities in the map (Fig. 23). The Code is used as the locality code and the colors represent the population cluster in the phylogenetic trees presented in the results (Fig. 16)...... 59 Table 8: Output of iPyRAD for number of total filtered loci and SNPs including and not including outgroups...... 63 Table 9: Results of model selection using AIC for the six demographic models analyzed (Figure 6), the models in bold were the selected for the bootstrap analyses...... 73 Table 10: Results of the f3-statistic for all species...... 73 Table 11: Population genetic diversity of each species, sample size (n), inbreeding coefficient (Fis) and average expected heterozygosity (Hs)...... 74 Table 12: Pair-wise Fst values between populations for each species...... 74 Table 13: Analysis of molecular variance for each species, testing for differences in geomorphologic processes (geo) and climatic stability (climatic). The values in bold represent significant variation (P ³ 0.001)...... 78

xii

LIST OF FIGURES

Figure 1. South America vegetation domains (source: The Nature Conservancy)...... 17 Figure 2. Brazilian biomes (source: Instituto Brasileiro de Geografia e Estatística - IBGE)...... 18 Figure 3. Chiasmocleis albopunctata from Catalão, Goiás, Brazil. Photo: Daniel Velho...... 19 Figure 4. Dendropsophus rubicundulus from Brasília, Distrito Federal, Brazil. Photo: Ísis da Costa Arantes...... 20 Figure 5. Physalaemus nattereri. Photo: Carlos Cândido...... 21 Figure 6. Maximum likelihood tree for Chiasmocleis albopunctata species group obtained from RAxML (20% missing data). The localities are represented by the Code (Table 2) and species are represented by the colors: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange = C. centralis...... 43 Figure 7. A) Population assignment results from processing of STRUCTURE analyses using the Evano DK method (Evanno et al., 2005) to identify the number of distinct genetic groups within the sample of Chiasmocleis; B) STRUCTURE bar plot illustrating probability of assignment to the six genetic clusters found for the Chiasmocleis albopunctata species group. Species are represented by the colors: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange and brown = C. centralis...... 44 Figure 8. Coalescent species tree of the Chiasmocleis albopunctata species group obtained from SVDquartets...... 45 Figure 9. Coalescent species tree of the Chiasmocleis albopunctata species group obtained from SVDquartets (Table 2). Species are represented by the colors: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange = C. centralis...... 46 Figure 10. Distribution map of the member of the Chiasmocleis albopunctata species group in which sampling localities for each species are represented by the colored circles: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange = C. centralis; red outline delimits the boundaries of the Cerrado biome; numbers correspond to localities listed in Table 2; the gray scale is elevation (m)...... 47 Figure 11. Coalescent model species tree for the Chiasmocleis albopunctata species group with divergence time estimates obtained from SNAPP; horizontal gray bars represent the 95% credibility interval of divergence time estimates at each node; numbers above branches indicate posterior probability for monophyly of the descendant clade. The x-axis represents the time before present (Mya)...... 48 Figure 12. Species tree obtained from species delimitation analysis for Chiasmocleis albopunctata species complex and the three species model that lumps C. centralis, C. bicegoi and C. sp into one species. Posterior probability is shown on the branch...... 49 Figure 13. Demography models tested using momi2. (1) divergence with no migration not considering the population size; (2) divergence with no migration considering the estimated population size; (3) divergence with bidirectional migration between populations A and B, considering the estimated population size; (4) divergence with bidirectional migration between populations B and C, considering the estimated population size; (5) divergence with unidirectional migration between populations C and B, and B and A, considering the estimated population size; (6) divergence with unidirectional migration between populations B and A, B and C, considering the estimated population size...... 85 Figure 14. Maximum likelihood tree for Chiasmocleis centralis obtained from RAxML. Populations are represented by the colors: purple = West; orange = Central-East; and blue = Central-North...... 86 Figure 15. Maximum likelihood tree for obtained from RAxML. Populations are represented by the colors: purple = South; orange = Central; and blue = North...... 87 Figure 16. Maximum likelihood tree for Physalaemus nattereri obtained from RAxML. Populations are represented by the colors: purple = West; orange = Central; and blue = North...... 88

xiii

Figure 17. Coalescent model species tree for Chiasmocleis centralis obtained from SVDquartets. Populations are represented by the colors: purple = West; orange = Central-East; and blue = Central-North...... 89 Figure 18. Coalescent model species tree for Dendropsophus minutus obtained from SVDquartets. Populations are represented by the colors: purple = South; orange = Central; and blue = North...... 90 Figure 19. Coalescent model species tree for Physalaemus nattereri obtained from SVDquartets. Populations are represented by the colors: purple = West; orange = Central; and blue = North...... 91 Figure 20. The three genetic clusters found, mean likelihood and DK for (A, b and c) Dendropsophus rubicundulus, (D, c and e) Chiasmocleis centralis, and (G, h and i) Physalaemus nattereri...... 92 Figure 21. Distribution map of Chiasmocleis centralis, in red the Cerrado range and number are the localities (Table 3), the gray scale is elevation (m)...... 93 Figure 22. Distribution map of Dendropsophus rubicundulus, in red the Cerrado range and number are the localities (Table 4), the gray scale is elevation (m)...... 94 Figure 23. Distribution map of Physalaemus nattereri, in red the Cerrado range and number are the localities (Table 5), the gray scale is elevation (m)...... 95 Figure 24. The demographic models selected for (A) Chiasmocleis centralis, (B) Dendropsophus rubicundulus, and (C) Physalaemus nattereri...... 96 Figure 25. Species distribution modeling for Chiasmocleis centralis under current climatic condition and projected for 12 different past climate data using Maxent...... 97 Figure 26. Climatic stability surface using all time periods for Chiasmocleis centralis, values ranging from 0 (no present) to 12 (present in all time periods), areas with high climatic stability have high values...... 98 Figure 27. Species distribution modeling for Dendropsophus rubicundulus under current climatic condition and projected for 12 different past climate data using Maxent...... 99 Figure 28. Climatic stability surface using all time periods for Dendropsophus rubicundulus, values ranging from 0 (no present) to 12 (present in all time periods), areas with high climatic stability have high values...... 100 Figure 29. Species distribution modeling for Physalaemus nattereri under current climatic condition and projected for 12 different past climate data using Maxent...... 101 Figure 30. Climatic stability surface using all time periods for Physalaemus nattereri, values ranging from 0 (no present) to 12 (present in all time periods), areas with high climatic stability have high values...... 102

xiv

GENERAL INTRODUCTION

Since the mid-nineteenth century, scientists such as Darwin, Hooker, Wallace and Sclater have sought to understand the distribution of species through evolutionary processes (Humphries

& Parenti, 1999; Brown & Lomolino, 2006). Knowledge of geographical distributions and their determinants is essential for understanding ecological and evolutionary aspects of organisms biology (Wiens & Graham, 2005; Morrone, 2009). The Neotropical region is one of the world's most biodiverse regions and has a great diversity of (Fig. 1) and a complex geological history, which have undoubtedly contributed to the distribution patterns of the native and plants (Morrone, 2014).

Though widely recognized for the characteristic rain that cover much of the

Amazon Basin, Andes, and , the Neotropics are also home to a number of unique and biodiverse savannas. The Neotropical savannas known as Cerrado, the largest savanna biome of the New World, and are mainly distributed in the Brazilian Shield (Fig. 2) and have a complex geological and evolutionary history, which together have shaped their current regional setting and biodiversity. The Cerrado is part of the SW-NE diagonal of open-biomes in South America, which also include Seasonally-Dry Tropical Forests (Caatinga) and Chaco (Werneck, 2011). The

South American open vegetations were formed by vicariant events of the Cenozoic (Paleogene and Neogene) period (Morrone, 2014) and each biome (Cerrado, Caatinga and Chaco) are

1

considered distinct vegetation domains (Werneck, 2011). The sequence of events that influenced the geographic extent and the biotic composition of these open vegetation regions throughout the

Cenozoic had a significant impact on the fauna of the region. The importance of these events to diversification of the Neotropical biota has been addressed in studies of (Giugliano et al.,

2007), amphibians (Santos et al., 2009; Maciel et al., 2010), birds (Brumfield & Edwards, 2007) and mammals (Hoffmann & Baker, 2003), which have revealed the Cerrado and Chaco to be more similar than either is to the Caatinga (Werneck, 2011). While these studies have largely focused on continental scale patterns of biodiversity, an understanding of patterns within these dry biomes and how local geography and climatic history may have influenced patterns of biodiversity within each is lacking.

The Cerrado: Regional setting

The Cerrado savanna is located in the core of South America (Fig. 2), on the Brazilian

Shield, occurring in Brazil, and (Ab’Saber, 1977) and is characterized by xeric vegetation ranging from savannas to forest formations (Eiten, 1972). The typical vegetation of

Cerrado varies from a dense with sparse shrubs and small trees to riverine forests, an almost closed woodland known as gallery forests, or other moist vegetation following the watercourses in the lower lands of the Brazilian Plateau (Ratter et al., 1997; Oliveira-Filho &

Ratter, 2002). Though present classification schemes limit the geographic extent of the biome to

Brazil, a recent study corroborated with Ab’Saber (1977) and suggested the Bolivian savannas should be incorporated in the Cerrado range (Werneck et al., 2012b). The Cerrado boundary

2

interdigitates with, and is influenced by other adjacent biomes (i.e., Amazon, Atlantic Forest,

Caatinga, Chaco, Pantanal) (Motta et al., 2002; Giselda, 2006), having isolated enclaves in these regions that are considered remnants from a period of greater extent for the biome, expansion of which was driven by climatic fluctuations (Eiten, 1972; Werneck et al., 2012b). Climate in this region is characterized by two well-defined seasons: a dry period (from May to September) and a wet period (from October to April) (Nimer, 1989). Annual average precipitation (1250-1500 mm) and temperature averages (20-22 °C) (Nimer, 1989), make for a relatively cool and dry environment. The soil is characterized by it is a lateritic soil, depleted in calcium, sodium, and magnesium (Oliveira-Filho & Ratter, 2002). It is the interactions between these climatic, topologic, and edaphic factors that presently limit the Cerrado to the central Brazilian plateau and Bolivia (Oliveira & Marquis, 2002).

Biodiversity and conservation status

The Cerrado is the second largest biome in South America and, before the recent colonization by European descendants, covered 22% of the country of Brazil, or about 1.7 million km2 (Myers et al., 2000; Jepson, 2005). However, 80% of its original area has been devastated for agropastoral activities and it is expected that the remaining will continue to decline in the near future (Myers et al., 2000; Jepson, 2005). The devastation of the Cerrado has been catastrophic, especially in the last three decades (Jepson, 2005). The rate of deforestation in the Cerrado is higher than that of the Amazon, with losses ranging from 22,000 to 30,000 km2 per year (Klink & Machado, 2005), and it is estimated that in a few years the original vegetation will be limited entirely to areas lying within currently established protected areas (Conservation

3

Units). The Brazilian Conservation Units (CUs) include two groups of protected areas: Integral

Protection Conservation Units (IPCUs), and Sustainable Use Conservation Units (SUCUs). In

IPCUs, the aim is to preserve areas important for biodiversity conservation by only permitting the indirect use of natural resources. In SUCUs, on the other hand, the sustainable use of an area’s natural resources is allowed. It is less restrictive than for IPCUs, and permits local and indigenous communities to sustainably use their lands.

The Cerrado has high species diversity and a large number of endemic species, and is considered the most diverse savanna in the world (Klink & Machado, 2005). Among the

Brazilian biomes, Cerrado has the greatest percentage of endemic amphibians, and plants,

51.7% of the 209 known amphibian species (Valdujo et al., 2012), 39% of the 267 squamate species (76 lizards, 158 snakes, and 33 amphisbaenians) (Nogueira et al., 2011), and 44% of the

7,000 plant species are endemic (Klink & Machado, 2005). The Cerrado biome is considered a global hotspot of biodiversity because of its diversity (Myers et al., 2000; Klink & Machado,

2005; Rylands & Brandon, 2005; Salgado & Galinkin, 2008), and due to the high level of human occupation and alteration it is a priority on the global agenda of conservation action. Given the enormous loss of natural areas within Brazil, CUs are essential for maintenance of remnants of

Brazilian biodiversity (Rylands & Brandon, 2005).

Geomorphologic events and climatic fluctuation

The diversification of the Neotropical biota occurred over millions of years in the face of dramatic geomorphological and climatic fluctuations. Two events that shaped the extant faunal biodiversity in South America altered the continent’s physical connection with other landmasses.

4

One event began in the Mesozoic, where the first notable event was the complete separation between South America and Africa during the Upper Jurassic around 160 million years ago

(Mya), putting an end to faunal exchange between the two continents (Bigarella & Andrade-

Lima, 1982). During the Paleocene (~66 – 56 Mya), the biotic exchange between North America and South America was interrupted by the disappearance of the land bridge connecting them, and

South America remained completely isolated until their reconnection in the Pliocene (~5 - 2.5

Mya) with the formation of the isthmus of Panama, reestablishing a terrestrial route for biotic exchange (Bigarella & Andrade-Lima, 1982).

The Brazilian and the Guiana Shields are ancient geologic formations, existent since before the Paleozoic (320 – 250 Mya). Their structure was altered by further upwarping during the Cretaceous (100 – 66 Mya) with the coincident formation of an erosion surface and the resulting subsidence of the Amazon Basin (Haffer, 1974). During the Miocene (20 - 5 Mya), marine incurssion occurred, surrounding the Guiana Shield and Brazilian Shield, separating these uplifted areas from one another and the recently risen Andes mountains to the west. This introgression divided the cis-Andean South American region into north and the south regions

(Morrone, 2014). There is clear evidence of the presence of marine introgression (Paranaense) in the Chaco-Pampean area during the middle and late Miocene (~ 13 - 5 Mya) (Cione et al., 2011).

In addition, periods of colder climate during this period were associated with the lowering of sea level, leading to the formation of shallow inland lakes, becoming a favorable environment for dispersal of terrestrial organisms among regions previously isolated by inland seaways (Haffer,

1979).

The geological events of the Neogene that had the great impact on the biota’s dispersal and speciation was the elevation of the Andes that began in the Cretaceous, became more active

5

in the Miocene (23-7 Mya), and culminated with the late uplift of northern Andes 5 - 2 Mya

(Haffer, 1974; Morrone, 2014). The late rise of the Andes allowed the immediate surroundings that were below the sea level through the Pliocene to finally emerged, as well as to the establishment of eastward drainage for all waterways in what is now known as the Amazon basin

(Haffer, 1974), many of which previously drained into the Pacific or north, through , to the Caribbean Sea. The division between south-western Amazonia and north-western South

America-Mesoamerica by the marine introgression through the Maracaibo and Amazon basins in the Pliocene (5 - 2 Mya) (Morrone, 2014) contributed to the formation of large South American rivers and the diversification of the Neotropical fauna (Rull, 2008; Vallinoto et al., 2010), especially in the central Brazilian plateau (Colli, 2005).

Around the same time as the final uplift of the Andes, the most important event for the diversification of Cerrado biota was the epeirogenic uplift of the central Brazilian plateau during the Late Miocene and Early Pliocene (~5 Mya), and subsidence of peripheral depressions

(Cerrado and Chaco) (Mendes & Petri, 1971; Haffer, 1974). This geologic/geographic transition among adjacent regions promoted speciation across elevation gradients (plateaus vs. valleys)

(Mendes & Petri, 1971; Haffer, 1974; Colli, 2005). These relatively recent geologic events in the

Cerrado were temporally coincident with climatic fluctuations during Quaternary, themselves essential to the diversification of the Cerrado biota (Prado et al., 2012; Werneck et al., 2012b).

The Cerrado savannas of the Last Glacial Maximum (LGM; 21 thousand years ago, or 21 Kya) experienced a retraction of their distributions with their geographic extent limited to the north- eastern portion of what is today the core region of this biome, known as the Serra Geral de Goiás refugium, and a number of fragmented, Amazonian enclaves (Werneck et al., 2012b). During the

LGM the climate was, relative to today, dry and cold. At this time, the Cerrado was distributed

6

on the plateaus and the peripheral depressions were dominated by xeric vegetation characteristic of today's Caatinga and Chaco (Werneck, 2011; Werneck et al., 2012b). During the mid-

Holocene (6 Kya) the climate became very similar to the current climate, semihumid warm tropical, contributing to the development of moist forests throughout Amazonia and a more diverse Cerrado vegetation (Behling, 1995; Giselda, 2006). During this time the savannas expanded and reached a distribution similar to their present-day distribution (Werneck et al.,

2012b). Notably, areas of climatic stability during Quaternary climatic fluctuations were important biotic refuges and may serve as cores of biome species richness (Werneck et al.,

2012b) and genetic diversity (Carnaval et al., 2009).

Several hypotheses have been proposed that explicitly consider the implications of geomorphological compartmentalization of the Cerrado landscape on the phylogenetic and phylogeographic patterns of the floral and faunal species inhabiting the region (Werneck, 2011).

Similarly, the expansion and contraction of forests in response to climatic changes and the consequent effects on the demography of Cerrado species (Rodrigues, 2005) as well as the climate-driven fragmentation of Cerrado savanna habitat all may have contributed to patterns of speciation (Vanzolini & Williams, 1981). Among these hypotheses are the Plateau/Valley hypothesis, the Gallery Forest hypothesis, and the Vanishing Refuge Model, respectively (Table

1).

The Plateau/Valley hypothesis was proposed on the basis of observations of phylogenetic and phylogeographic patterns of species limited in their distributions to the Cerrado (Werneck,

2011). This hypothesis predicts that reciprocally monophyletic groups of species associated with savanna and forest formations are expected to be associated with plateaus and valleys, respectively (Werneck, 2011). In addition, plateau regions are more ancient and are therefore

7

expected to have high genealogical structure and genetic diversity with evidence of recent expansion into Cerrado valleys, where less genetic diversity is expected (Werneck, 2011). The gallery forests are thought to have maintained the connectivity between the central Brazilian plateau and adjacent forests during the least favorable climatic periods, providing corridors and refuge to a stable forest biota (Brown & Gifford, 2002). This hypothesis was tested in a study of the lizards Micrablepharus atticolus, which is a Cerrado endemic, however there were no differences in genetic diversity between plateau vs. valleys and stable vs. unstable areas (Santos et al., 2014).

The Gallery Forest hypothesis relies on these forested areas as modulators of speciation; during humid periods when the forests are expected to have expanded their distributions, they would have encroached upon and fragmented open areas. Under such a scenario, the gallery forests might have provided shelter for faunal groups otherwise adapted to open formations

(Rodrigues, 2005). The open-habitat species might have had minimal pre-adaptations to forested habitats, and they could freely visit and tolerate gallery forest; however, gallery forest species could not tolerate open habitats and they were strictly ombrophilous (Rodrigues, 2005). During these humid periods characterized by expansion of forests, the species adapted to open vegetation became isolated from the core of the Cerrado and were confined within the gallery forest, promoting allopatric speciation (Rodrigues, 2005). Thus, the gallery forest refuges contribute asymmetrically to the diversity of rainforest species than to the diversity of the

Cerrado (Rodrigues, 2005). On the other hand, small-bodied species with low vagility (i.e. lizards and amphibians) might have been less affected by these Quaternary climatic fluctuations, having higher richness and endemicity in the open habitats of plateaus than the gallery forest of adjacent depressions (Nogueira et al., 2009). This hypothesis has yet to be rigorously tested, but

8

a phylogenetic study of lizards of the Kentropyx that possesses both savanna and forest specialists demonstrated the monophyly of the forest-inhabiting species, ruling out the savanna- inhabiting species as a source for gallery forest species as the hypothesis proposes (Werneck et al., 2009).

The Vanishing Refuge Model (VRM) is in some ways the inverse of the Gallery Forest hypothesis, proposing that populations of forest-restricted species may have been pre-adapted to open habitats and during dry periods of climatic fluctuations, they became confined to a forested refuge that eventually vanishes (Vanzolini & Williams, 1981). In this process the species may become completely adapted to open formation and constitute a full ecological variant (Vanzolini

& Williams, 1981). A recent study tested this model and developed genetic predictions for species derived from VRM processes (Damasceno et al., 2014). The VRM thus proposes climate-driven habitat fragmentation of the forest promotes genetic isolation of forest-restricted species, which combined with eventual adaptation to an open habitat, leads to allopatric speciation (Vanzolini & Williams, 1981; Damasceno et al., 2014). In the initial stages of VRM, the model predicts the disappearance of habitat connections between isolated patches of forest and genetic differentiation between populations in core areas and those inhabiting vanishing refugia (Damasceno et al., 2014). After that, it is expected that distinct genetic lineages will persist in disjunct habitat patches (core and vanishing refugia), followed by signs of population expansion as isolates adapt to new (open) habitats, differentiate phenotypically, and eventually evolve reproductive isolation (Damasceno et al., 2014). Support for the recently-proposed VRM model has been provided by Damasceno et al. (2014), where the shallow phylogeographic structure and ecological differentiation among isolates of the Atlantic-Forest-dwelling

9

species Coleodactylus meridionalis is interpreted as suggesting that these populations are in the earlier stages of diversification under the VRM.

These hypotheses have only recently been applied to the exploration of patterns of diversity in the Neotropical savannas. I have used these hypotheses as a guide in the development of my research in which I explore the biogographic and phylogeographic patterns of anurans (frogs) in the Cerrado. My goal is to use anurans as a model with which to test these hypotheses in order to better understand the evolution of the biota of the Neotropical savannas.

10

Table 1: Hypotheses of Cerrado diversification and their predicted effects on the evolutionary history of herpetofauna. Hypothesis Timing Factors Effects Predictions Plateau/Valley Neogene Varying Plateau regions are more Plateaus are expected to harbor (Werneck, environmental ancient, recent expansion high genealogical structure and 2011) conditions of into Cerrado valleys. genetic diversity, with recent plateaus vs. expansion into forested valleys valleys over time. evidenced by less genetic diversity in valleys. Reciprocally monophyletic groups of species associated with savanna and forest formations. Gallery Forest Quaternary Humid periods The Gallery Forest Gallery Forest refuges (Rodrigues, leading to provided shelter for contribute asymmetrically with 2005) expansions of faunal groups adapted to more species to the rainforest rainforest over the open formations. than to the Cerrado. Higher open areas and Then, the species richness of species in Gallery encompassing become isolated from the Forest than open vegetation is Gallery Forests. Cerrado and were expected. Polyphyletic groups confined in the Gallery of species associated with Forest, promoting Gallery Forest. allopatric speciation. Vanishing Quaternary Climatic-driven Populations of forest- Recent habitat connections Refuge Model habitat restricted species may be between isolated patches and (Vanzolini & fragmentation and pre-adapted to open genetic differentiation between Williams, 1981; refuge vanishing. habitats and during dry populations in core areas and Damasceno et periods they are confined vanishing refugia. After that, it al., 2014) to a refuge that is expected to find distinct eventually vanishes. lineages in different habitats During the gradual (core vs. vanishing refugia), erosion of their refuge, signs of population expansion species may become in new habitats, phenotypic adapted to open differentiation, and formation and constitute reproductive isolation. a full ecological variant.

11

Biogeography of Cerrado Amphibians

The geomorphological changes in the Neogene (Miocene - Pliocene) and Pleistocene climatic fluctuations are widely recognized as driving forces for the diversification of extant amphibians in South America. These events had a strong impact on diversification within the region and several studies have demonstrated concordant patterns with these events in different species (Santos et al., 2009; Vallinoto et al., 2010; Gehara et al., 2014). The biogeography of some South American frogs shows concordant diversification patterns, including the poison frogs (Dendrobatidae) (Santos et al., 2009), Rhinella marina group (Maciel et al., 2010;

Vallinoto et al., 2010), Dendropsophus minutus (Gehara et al., 2014) and Hypsiboas albopunctatus (Prado et al., 2012). The major geological events that influenced the diversification of South American anurans were the uplift of the Andes and the Amazonian basin formation during the Miocene – Pliocene (Santos et al., 2009; Maciel et al., 2010; Gehara et al.,

2014). The geomorphological changes described above certainly influenced the continuity of species ranges and thus, influenced patterns of speciation, as has been demonstrated in taxa such as the Rhinella marina group and Hypsiboas albopunctatus (Maciel et al., 2010; Prado et al.,

2012). Furthermore, recent studies have clearly demonstrated the climatically-influenced intermittent biogeographic connections between the moist Amazon and Atlantic Forests, with alternating connections between the regions open/xeric biomes (Fouquet et al., 2014; Gehara et al., 2014; Ledo & Colli, 2017; de Sá et al., 2019). The Cerrado biome was clearly influenced by these same climatic and geomorphological processes, experiencing both contractions and expansions, all while being primarily limited to the Brazilian central plateaus (Werneck et al.,

2012b). This localization of the Cerrado helps explain the apparent pattern of in situ

12

diversification that characterizes the biota of the Brazilian Central Shield (Maciel et al., 2010;

Prado et al., 2012) and highlights the importance of the regional-scale hypotheses described above for explaining this diversification.

To date, few studies have focused on diversification of the amphibians of the Cerrado savannas, and most of the previous works addressed this indirectly as they were focused primarily on the drivers of high tropical diversity in the forested biomes. In this study I seek to reveal general patterns of amphibian diversification in the Cerrado, and to identify of events that commonly affected their diversification and presumably contributed to the origin of the high diversity in this unique region.

Anurans as a Model of Cerrado Diversification

The three focal species of this study (Chiasmocleis albopunctata, Dendropsophus rubicundulus and Physalaemus nattereri) were chosen because of their broad distribution and near endemicity to the Cerrado savannas (Valdujo et al., 2012). In addition, these species

(Chiasmocleis albopunctata, Dendropsophus rubicundulus and Physalaemus nattereri) belong to different families and have dramatically different morphology and life-history strategies; however, they have a similar distribution in the Cerrado and all are associated with open habitats.

The white-spotted humming frog, Chiasmocleis albopunctata (Microhylidae; Boettger,

1985) (Fig. 3), is distributed in the savannas of (Sánchez et al., 2014), Bolivia (De La

Riva et al., 1996), Brazil (Western, Southeastern and Northern), and Paraguay (Forlani et al.,

2011). The species’ breeding behavior occurs during the rainy season of December through

March in temporary water bodies, wetlands and (Aquino et al., 2004a; Silva et al.,

13

2009; Vasconcelos et al., 2011). The color pattern of this species consists of a dark gray or dark brown dorsum, legs and arms with white blotches, a whitish bar on the snout, extending through the canthus rostralis and upper eyelid, and the belly is brownish or grayish with irregular whitish blotches. The males’ calls varies geographically, the Brazilian populations have a call characterized by notes with nine pulses and two harmonic frequency bands (Oliveira-Filho &

Giaretta, 2006), whereas the Bolivian populations have a call with five to eight pulses and no harmonics (De La Riva et al., 1996), suggesting this may represent a species complex (Oliveira-

Filho & Giaretta, 2006).

The Lagoa Santa treefrog, Dendropsophus rubicundulus (Hylidae; Reinhardt and Lütken,

1862) (Fig. 4), occurs in the Cerrado’s savannas of Brazil (Southeastern, Northeastern,

Northern), and similarly open habitats in adjacent regions of Bolivia, and Paraguay. This species has the widest distribution within the D. rubicundulus species group, which comprises: D. rubicundulus; D. tritaeniatus; D. anataliasiasi; D. araguaya; D. cerradensis; D. cachimbo; D. elianeae; D. jimi; and D. rhea (Napoli & Caramaschi, 1999; 2000; Faivovich et al., 2005;

Brusquetti & Lavilla, 2006; Annunziata et al., 2007; Silva et al., 2011; Frost, 2017). The color patterns are variable, although the predominant pattern in life is green and smooth dorsum with a lateral darker band from the snout tip to nearly thighs that separates the lighter color in the dorsum (Reinhardt & Lütken, 1862; Lutz, 1973). Breeding occurs during the rainy season, from

October to January (Moreira & Barreto, 1997), and the calling sites are in emergent and marginal shrubs near to permanent or temporary ponds, pools, and lagoons in open areas or grasslands

(Cardoso & Vielliard, 1984; Colli et al., 2004). Like C. albopunctata, it has been suggested that this species, as currently defined, might be a species complex in need of taxonomic review and revision (Colli et al., 2004).

14

The Cuyaba dwarf frog, Physalaemus nattereri (Leptodactylidae; Steindachner, 1863)

(Fig. 5), is distributed in open areas in Brazil (Central and Southeastern) (Aquino et al., 2004b;

Frost, 2017), Bolivia (De La Riva et al., 2000), and Paraguay (Brusquetti & Lavilla, 2006; Smith et al., 2012). The coloration of the dorsum is reddish brown, regularly scattered with dark and light brown spots, and two evident black inguinal glands (black ocelli) (Cei, 1980). The reproductive period occurs mostly during the rainy season, from September to January

(Brasileiro et al., 2005), in temporary and permanent water bodies in open and disturbed areas

(pasture fields and anthropic forest fragments) (Brasileiro et al., 2005; 2008; Giaretta & Facure,

2006; Santos et al., 2007; Affonso et al., 2014). Unlike C. albopunctata and D. rubricundulus, there has been no suggestion in the literature that this species represents a complex of undescribed/unrecognized species.

The three species are listed as IUCN Least Concern because of their wide distributions and large populations (Aquino et al., 2004a; b; Colli et al., 2004). Of these three species, only

Chiasmocleis albopunctata is tolerant to a large range of habitats and seems to be tolerant of anthropogenic disturbance (Aquino et al., 2004a). Populations of Dendropsophus rubicundulus and Physalaemus nattereri are decreasing in size and number, indicating they are not tolerant to anthropogenic disturbance (Aquino et al., 2004b; Colli et al., 2004). The taxonomic and life- history diversity of these three species make them suitable focal species for investigating the history of biotic diversification in the Cerrado.

Objectives

15

The aim this study is to reveal the evolutionary processes that contributed to diversification in the Cerrado savanna using amphibians as a model. Evolutionary patterns were investigated by comparing the phylogeographic and biogeographic history of biome-specific yet widely distributed amphibian species (Chiasmocleis albopunctata, Dendropsophus rubicundulus and Physalaemus nattereri). Congruence in phylogeographic patterns of these taxa that vary so dramatically in their ecologies would suggest that lineages within the region have responded similarly to environmental changes. The synthesis of observed patterns of evolutionary diversification within the region will enable the testing of hypotheses proposed as general patterns for diversification in Cerrado, as well as the identification of common events that affected the diversification of my focal taxa. Understanding the species diversification process and the response of savanna-inhabiting anuran species to past and ongoing environmental changes is critical to accurate characterization of the tropical speciation process and adequate conservation of this unique region.

16

FIGURES

Figure 1. South America vegetation domains (source: The Nature Conservancy).

17

Figure 2. Brazilian biomes (source: Instituto Brasileiro de Geografia e Estatística - IBGE).

18

Figure 3. Chiasmocleis albopunctata from Catalão, Goiás, Brazil. Photo: Daniel Velho.

19

Figure 4. Dendropsophus rubicundulus from Brasília, Distrito Federal, Brazil. Photo: Ísis da Costa Arantes.

20

Figure 5. Physalaemus nattereri. Photo: Carlos Cândido.

21

CHAPTER 1

SPECIES DELIMITATION OF CHIASMOCLEIS ALBOPUNTATA SPECIES COMPLEX

INTRODUCTION

The rain forests of the Neotropics are widely recognized for their great biological diversity. However, it is less widely known that the Neotropical savannas have exceptionally high levels of species diversity and endemism, and are in fact considered the most diverse savannas in the world (Klink & Machado, 2005). The level of endemicity is elevated in the

Cerrado biome, the largest of the Neotropical savannas, with, for example, nearly 52% of the amphibian species known to be endemic (Valdujo et al., 2012). This great diversity/endemicity, combined with the high levels of deforestation, make this biome a global hotspot of biodiversity

(Myers et al., 2000; Klink & Machado, 2005; Rylands & Brandon, 2005; Salgado & Galinkin,

2008). A lot of the Cerrado diversity has yet to be discovered, and the taxonomic, distributional, and phylogenetic knowledge gaps are difficult to overcome. This is due, in equal parts, to three shortfalls that afflict our knowledge of biodiversity: Linnean (lack of taxonomic knowledge),

Wallacean (lack of distributional knowledge), and Darwinian (lack of phylogenetic

22

context/understanding). These shortfalls are very accentuated in the Neotropics due, in part, to the extreme isolation of some areas (Domingos et al., 2017) and the overwhelming scope of the biota in question. Conversely, with the advent of genetic methods, more and more studies are revealing new species (Vaz-Silva & Maciel, 2011; Fouquet et al., 2014; Haga et al., 2017) or highlighting the presence of cryptic species (Domingos et al., 2014; 2017; Gehara et al., 2014;

Guarnizo et al., 2016), ever increasing the number of species in the Neotropical region.

Next-generation sequencing is a relatively new methodology and has been used in several studies of population genetics, phylogeography, species delimitation, and high-resolution genetic mapping (Baird et al., 2008; Davey et al., 2010; Emerson et al., 2010). This method is an inexpensive, fast and accurate way to obtain genome-wide information (Peterson et al., 2012) and it does not require the prior development of any genomic resources (Emerson et al., 2010).

However, the amount of data generated with next-generation sequencing is enormous, making data analysis a dauting task that requires knowledge in bioinfomatics, access to a large number of computer cores and memory, and large amounts of time to run appropriate analyses (McCormack et al., 2013). In the last decade, phylogenetic and phylogeographic studies have become increasingly reliant upon genomic-scale data, and consequently the development of new data analysis methods are improving outcomes and reducing the time required. Today, the most common approach to balance the ability to generate massive amounts of genetic data and analyze them effectively is to use a subset of genomic markers randomly distributed throughout the genome, obtaining a highly informative single nucleotide polymorphism (SNPs) dataset

(Amirisetty et al., 2012; Henriques et al., 2018). Recently, new methods for estimating gene trees without gene trees (i.e. coalescent methods to estimate speccies tree) are making possible the use of such genome-wide data for application to hypotheses of species delimitation (Bryant et al.,

23

2012) and have demonstrated that coalescent-based methods are robust tools for this purpose

(Fujita et al., 2012; Leaché et al., 2014). There have been few studies to apply such data and analytical methods to the biota of the Neotropical region (Gamble et al., 2012; Domingos et al.,

2014; 2017; Gehara et al., 2014), but those that have been published have revealed high levels of cryptic diversity for vertebrate species in the region.

Chiasmocleis albopunctata (Microhylidae; Boettger, 1985) is widely distributed in the

Cerrado and Chaco, occurring in Argentina (Sánchez et al., 2014), Bolivia (De La Riva et al.,

1996; Jansen et al., 2011), Brazil and Paraguay (Forlani et al., 2011). Due to its broad altitudinal and latitudinal geographic distribution, this species exhibits extensive morphological variation, and mating calls, indicating the likely presence of cryptic species. Recent studies of

Chiasmocleis that included morphological and genetic data have brought to light the obvious problems that exist within the current framework of species boundaries and phylogeographic patterns of the populations currently recognized as C. albopunctata. Forlani et al. (2011), reassessed the distribution of this species and suggested there are might be more than one species because the populations from Bolivia have a different call structure when compared with

Brazilian populations (De La Riva et al., 1996; Oliveira-Filho & Giaretta, 2006). A taxonomic revision and phylogeny of Chiasmocleis and Syncope revealed that C. albopunctata was nested within the Atlantic Forest species, and C. leucosticta was the sister species (Peloso et al., 2014).

A recent study using morphological and genetic data has also shown that origin of the C. albopunctata clade (C. albopunctata, C. mehelyi (Caramaschi & da Cruz, 1997), C. centralis

(Bokermann, 1952), and C. bicegoi (Miranda-Ribeiro, 1920) share a recent common ancestor with Southern Atlantic Forest clade that includes C. leucosticta, C. mantiqueira, and C. altomontana, diverging around 19 million years ago (Mya) (95% Highest Posterior Density

24

(HPD) = 14.03-24.24) (de Sá et al., 2019). In addition, the same study discovered that C. albopuncata is a sister taxon to the clade comprising C. mehelyi, C. centralis, and C. bicegoi, and also C. centralis and C. bicegoi are sister species to C. mehelyi taxon (de Sá et al., 2019).

Furthermore, the same study resurrected C. bicegoi as a valid species (de Sá et al., 2019), it was synonymy of C. albopunctata by Cruz et al. (1997). Though the knowledge of this species group has improved in the last decade, species boundaries, geographic distributions, and phylogeographic patterns are still in critical need of assessment.

Objective

Given the incredibly high levels of diversity in the Neotropical savannas, and the relatively little attention they have received from researchers exploring factors influencing patterns of biodiversity, this region is ripe for the testing of hypotheses proposed to explain evolutionary patterns. The Neotropical region are good environments to study patterns of diversification because of the elevated number of amphibian’s species. Particularly, the Cerrado savannas have high number of endemic and cryptic species, which makes this biome even more interest to study. The findings of recent studies have provided insight into the phylogenetic relationships and historical biogeography of the Chiasmocleis species distributed in the Cerrado and Chaco, though the granularity of these studies prevents the testing of speciation/diversification hypotheses that have been proposed to affect Neotropical species. The focus of this chapter is to explore patterns of biodiversity (including the identification of cryptic species) in the C. albopunctata species group. Whereas previous studies have employed few genetic makers, including mitochondrial and nuclear DNA, in combination with morphological

25

data, for this chapter I sampled thousands of loci randomly distributed throughout the genomes of specimens of Chiasmocleis. With these data I applied phylogenetic, phylogeographic, and coalescent species delimitation methods in order to resolve questions of evolutionary history, taxonomic diversity, species boundaries, and test hypotheses that implicate climatic and geomorphological events in the formation of this diversity.

Methods

Sampling

Fieldwork was conducted in December of 2016, in Brasília, Federal District, Brazil, I actively searched for and manually captured individuals of the Chiasmocleis albopunctata species group. Collected individuals were euthanized with lidocaine ointment or lethal injection of xylocaine. Tissue samples of muscle and/or liver were collected and preserved in cryogenic tubes with 95% ethanol until they could be placed in a -80º C freezer. Specimens were then fixed in 10% formalin and preserved in 70% ethanol. All specimens and tissues collected were deposited in the Coleção Herpetológica da Universidade de Brasília (CHUNB), Brasília – DF,

Brazil. Also, I gathered tissue samples of C. albopunctata from herpetological collections in

Brazil, Paraguay, Argentina and Germany.

For the study of phylogeographic patterns I gathered 197 tissue samples from 54 localities of Chiasmocleis albopunctata (Table 2). Furthermore, for phylogenetics analyses I used two samples of geayi, Dermatonotus muelleri, and Elachistocleis ovalis as outgroups.

26

Table 2: Sampling localities of Chiasmocleis albopunctata group specimens used in this study. The number (#) corresponds to the localities on the map (Fig. 10). The Code is used as the locality code in the phylogenetic trees presented in the results (Fig. 6). # Code Municipality Country State N ind Longitude Latitude 1 santc Santa Cruz Bolivia - 4 -62.0751 -16.7476 2 vilab Vila Bela da Santíssima Trindade Brazil MT 9 -60.0004 -14.0154 3 jauru Jaurú Brazil MT 7 -58.8623 -15.3174 4 tanga Tangará da Serra Brazil MT 3 -58.3230 -14.4314 5 lamba Lambari D'Oeste Brazil MT 1 -57.7469 -15.1599 6 cuiab Cuiabá Brazil MT 4 -55.8894 -15.4800 7 chapa Chapada dos Guimarães Brazil MT 9 -55.5399 -15.1084 8 novau Nova Ubiratã Brazil MT 1 -54.7664 -12.9931 9 tesou Tesouro Brazil MT 4 -53.4580 -15.8890 10 ribei Ribeirão Cascalheira Brazil MT 4 -52.0810 -12.8579 11 nossa Nossa Senhora do Livramento Brazil MT 1 -56.3747 -16.2399 12 pocon Poconé Brazil MT 3 -56.9485 -16.7845 13 corum Corumbá Brazil MS 2 -56.7238 -18.7311 14 jardi Jardim Brazil MS 1 -56.1494 -21.4804 15 amamb Amambay Paraguay - 4 -56.0259 -22.6539 16 centr Central Paraguay - 1 -57.5140 -25.1867 17 formo Mayor Villafañe, Formosa Argentina - 4 -59.0833 -26.2000 18 canin Canindeyú Paraguay - 1 -55.5283 -24.1348 19 assis Assis Brazil SP 4 -52.3780 -22.5902 20 teodo Teodoro Sampaio Brazil SP 6 -52.2975 -22.5290 21 tresl Três Lagoas Brazil MS 2 -51.7851 -21.0253 22 bauru Bauru Brazil SP 7 -49.0838 -22.2342 23 buri Buri Brazil SP 1 -48.5754 -23.7518 24 pauli Paulínia Brazil SP 2 -47.1449 -22.7478 25 cosmo Cosmópolis Brazil SP 1 -47.1011 -22.4812 26 conch Conchal Brazil SP 1 -47.1422 -22.3691 27 mogig Mogi Guaçú Brazil SP 2 -47.1747 -22.2781 28 riocl Rio Claro Brazil SP 1 -47.4667 -22.3667 29 pedre Pedregulho Brazil SP 2 -47.4361 -22.2247 30 itira Itirapina Brazil SP 3 -47.8891 -22.2175 31 luiza Luíz Antônio Brazil SP 19 -47.6188 -21.5866 32 conqu Conquista Brazil MG 1 -47.5419 -19.9369 33 sanri Santana do Riacho Brazil MG 3 -43.6039 -19.3435 34 curve Curvelo Brazil MG 8 -44.4487 -18.8207 35 burit Buritizeiro Brazil MG 3 -44.9619 -17.3508 36 sanfe Santa Fé Brazil MG 1 -45.4139 -16.6900 37 bonit Bonito de Minas Brazil MG 5 -44.7500 -15.2500 38 graom Grão Mogol Brazil MG 1 -42.5781 -16.7372 39 luzia Luziânia Brazil GO 5 -48.0251 -16.6749 40 brasi Brasília Brazil DF 8 -47.8761 -15.9396 41 nique Niquelândia Brazil GO 1 -48.4018 -14.4985 42 altop Alto Paraíso de Goiás Brazil GO 8 -47.5352 -14.1830 43 sdomi São Domingos Brazil GO 1 -46.4870 -13.5370 44 paran Paranã Brazil TO 6 -47.8122 -12.6959 45 peixe Peixe Brazil TO 6 -48.5552 -11.9968 46 palma Palmas Brazil TO 3 -48.1536 -10.2186 47 lajea Lajeado Brazil TO 4 -48.2899 -9.8483 48 casea Caseara Brazil TO 2 -49.9599 -9.3091 49 goiat Goiatins Brazil TO 3 -47.4811 -8.0936 50 palte Palmeirante Brazil TO 1 -48.1402 -7.8898 51 carol Carolina Brazil MA 8 -47.4136 -7.2980 52 estre Estreito Brazil MA 3 -47.2013 -6.7826 53 babac Babaculândia Brazil TO 1 -47.8559 -7.1751 54 palmt Palmeiras do Tocantins Brazil TO 1 -47.6425 -6.5997 N ind = number of individuals. MT= Mato Grosso; MS = Matogrosso do Sul; SP= São Paulo; MG = Minas Gerais; GO = Goiás; DF= Distrito Federal; TO = Tocantins; MA = Maranhão.

27

Molecular methods

Whole genomic DNA was extracted from tissue samples (muscle and/or liver) using the

DNA salt extraction protocol (Aljanabi & Martinez, 1997). In order to explore phylogeographic patterns I used the triple digest, restriction-site-associated DNA sequencing (3RADseq) method to generate sequence data from thousands of loci for all individuals (Glenn et al., 2017).

Advances in high-throughput sequencing have proven highly effective in a wide range of studies such as high-resolution genetic mapping (Baird et al., 2008) and population genetics (Davey et al., 2010), but has proven particularly effective in studies of phylogeography (Emerson et al.,

2010). This method has some advantages compared to previous methods: it does not require the prior development of any genomic resources (Emerson et al., 2010), it is an inexpensive and rapid method (Peterson et al., 2012), and it can be used for simple and complex analysis such as genotyping and single-nucleotide polymorphism (SNP) discovery, applicable to quantitative genetic and phylogeographic studies, respectively (Davey et al., 2010).

The 3RADseq method enables the identification, verification and scoring of thousands of

SNPs from anonymous regions in seven steps (Glenn et al., 2017):

1. Digestion – digestion of 100 ng of genomic DNA for 1 hour at 37º C with two restriction

endonucleases enzymes (XbaI and EcoRI) and a third enzyme (NheI), that digests the

dimers formed by the phosphorylation on the 5’ end of the iTru Read 1 adapter. These

enzymes were chosen based on an in silico digest of the genome of Xenopus laevis that

suggested the combination would yield an appropriate number of loci for downstream

analyses (i.e. not fewer than 2,000 loci, not more than 10,000 loci).

2. Ligation – DNA ligation with double-stranded adapters whose 'sticky' ends were

complimentary to those left by the restriction enzyme. The ligated adapters contain a

28

terminal region that is complimentary to primers that are used in step 4 below. Ligation

reactions were run using the following cycle: 22ºC for 20 min, 37ºC for 10 min, repeat

22°C and 37°C steps once, and 80ºC for 20 min).

3. Clean-up –SeraMag speedbeads were used, in combination with a 96-well plate magnet,

to purify the ligation reaction of any non-DNA molecules before next step. Ligation

reactions were washed twice with Ethanol and eluted in IDTE buffer.

4. Polymerase chain reaction (PCR) – All DNA fragments possessing an XbaI site on one

end and an EcoRI site on the other end were amplified using a combination of primers

(complimentary to the ligated adapter sequence) were amplified using iTru5 and iTru7

primers. These primers each possessed a unique barcode sequence that would enable

assignment of all reads to an individual frog. Each DNA sample was amplified with a

unique combination of iTru5/7 primer barcodes, such that each sample could be

identified. After amplification, PCR products were pooled in groups of approximately 24

samples with 5 to 10 ng per pool. Pools were then cleaned with the MinElute® PCR

Purification Kit (QIAGEN) to purify the PCR products.

5. Fragment size selection – 5 to 10 ng of the cleaned PCR product pool containing the

DNA fragments of interest were size separated on a Pippin PrepTM (Sage science) using

1.5% agarose gel cassettes. From this, fragments ranging in size from 328-418 base pairs

were separated and retained for sequencing.

6. Quantification of samples – Purified, size-selected fragment pools were quantified using

qPCR with the KAPA SYBR Fast qPCR kit on an Applied BiosystemsTM QuantStudioTM

real-time PCR system.

29

7. Sequencing – Pools were sequenced on Illumina NextSeq® 500/550 High Output Kit v2

(75 cycles) flow cells, at the National Center for Natural Products Research at the

University of Mississippi.

The DNA sequences obtained after the Illumina sequencing were processed using the following bioinformatic pipeline on a High-performance Computing Cluster at the Mississippi

Center for Supercomputing Research (MCSR). The raw sequences were demultiplexed by the individual barcodes and converted from .bcl to .fastq file format using bcl2fastq Conversion

Software available from Illumina (https://support.illumina.com/ sequencing/sequencing_software/bcl2fastq-conversion-software.html). The iPyRAD (v0.7.28) pipeline, a newer version of PyRAD, was used to perform de novo assembly of RADseq datasets

(Eaton, 2014), identifying loci within individuals and then creating alignments of loci across all individuals sequenced. The clustering similarity threshold was set to 0.85, and the offset for quality scores was left at the default value of 33, the minimum depth for statistical and the minimum depth majority rule for the consensus base calling was set to 10, and the maximum number of allowed barcode mismatches was set to 2 (our barcodes had, at minimum, 3 nucleotide degeneracy). For phylogenetic inference, I produced SNP datasets including the outgroups, and for population genetic inference, I produced SNP datasets without the outgroups.

To explore the effects of missing data, three different matrices were produced allowing 20%,

50%, and ~100% (i.e. no threshold was set) of missing data in the shared loci for all individuals

(Table 3).

30

Table 3: Output of iPyRAD for number of total filtered loci and SNPs including and not including outgroups. Species N Missing Min. ind. Loci Unlinked Total Outgroups data SNPs SNPs 20% 162 2,861 2,845 21,173 Yes 203 50% 101 8,137 8,031 53,511 Yes C. ~100% 4 51,451 36,978 165,008 Yes albopunctata 20% 157 3,216 3191 23,444 No 197 50% 98 8,423 8,302 54,510 No ~100% 4 51,301 36,771 162,886 No N= total samples; Min. ind. = minimal number of individuals per locus.

Phylogenetics, Phylogeography, and Population structure

I used two approaches to infer the phylogeographic structure from my SNP data. First, I

used RAxML v8.2 which performs a maximum likelihood-based inference of phylogenetic

relationships using SNP matrices with 20% missing data (Stamatakis, 2014). In addition, I used

Lewis (2001) acquisition bias correction, a conditional likelihood method for dealing with

variable sites. To remove the non-binary sites, the matrices with all SNPs were pre-processed

using the package Phrynomics in R (Leaché et al., 2015; R Development Core Team, 2017). I

used the K80 model of nucleotide substitution with ASC_GTRCAT rate heterogeneity. Branch

support for recovered phylogenetic groupings was estimated using an automatic bootstrap

function (autoMRE) (Pattengale et al., 2010). For the species tree inference using SNPs, I used

SVDquartets (Chifman & Kubatko, 2014) to apply a multi-species coalescent model

implemented in PAUP* v4.0 (Swofford, 2002). I used unlinked SNPs with 20% and 50$ of

missing data allowed, and with 12 million random quartets with 100 bootstrap replicates, setting

the outgroups as monophyletic groups.

To infer population genetic structure, I used unlinked SNP matrices with 20% of missing

data without outgroups. The analysis is a model-based clustering method performed in

STRUCTURE v2.3 (Pritchard et al., 2000), using the package ParallelStructure in R (Besnier &

31

Glover, 2013). First, I estimated the value of lambda to be used in the subsequent runs performing 10 runs with K=1. After that, I implemented 5 runs for each value of K ranging from

1 to 8 (40 runs) with 100,000 burn-in steps and a Markov chain Monte Carlo (MCMC) of

400,000 steps. I then ran the results of STRUCTURE through the STRUCTURE HARVESTER online software (Earl & vonHoldt, 2012), which explores how many distinct genetic groups there are within the data set and whether there is evidence for dispersal and/or gene flow. I then estimated the optimal k value by implementing the method of Evanno et al., (2005) and verified changes in the Ln P(D) of each model. I used CLUMPAK server to produce the graphics to illustrate

STRUCTURE results (Kopelman et al., 2015).

Divergence time

I used SNAPP (Bryant et al., 2012), a package available within the BEAST2 software package (Bouckaert et al., 2014), which uses SNP data to perform a MCMC-based Bayesian computation to estimate divergence times and effective population sizes. I used a reduced version of my sampling, including only one individual per locality and one sample of each outgroup included in the analysis, to reduce the time and computational requirements for this analysis. The SNP matrices used for this analysis had a limit of 50% of missing data and were filtered to only include bi-allelic SNP (1000 maximum) data (Ruby script available at https://raw.githubusercontent.com/mmatschiner/snapp_prep/master/snapp_prep.rb). I performed a time calibration of the tree using normal and lognormal distributions, and I used the time of most recent ancestor common ancestor (MRCA). I used 3 constraints: (1) the origin of

Gatrophryninae at 79.1 Mya using a normal distribution with 5 standard deviations (95%HPD

70.9-87.3 Mya) (de Sá et al., 2012); (2) the origin of Melanophryne + Ctenophryne +

32

Nelsonophryne clade was 21.85 Mya (95% HPD 11.00-37.36 Mya) according de Sá et al.

(2012), I used a lognormal prior with a mean of 21.85 Mya and a standard deviation of 0.35 Mya to allow a higher sampling probability around this value (95% HPD 11.6-36.5 Mya); (3) the most recent common ancestor of the southern Atlantic Forest clade (including Chiasmocleis leucosticta, C. mantiqueira, C. altomontana, and the Cerrado and Chaco species C. albopuntata,

C. mehelyi, C. centralis, and C. bicegoi), was 19 Mya (95% HPD 14.03-24.24 Mya) (de Sá et al.,

2019), I used a lognormal prior with a mean of 19.0 Mya and a standard deviation of 0.15 Mya

(95% HPD 14-24.2 Mya).

Species delimitation

The species delimitation method used was the Bayes factor delimitation (BDF*) (Leaché et al., 2014), using the packages SNAPP (Bryant et al., 2012) and Model selection (Baele et al.,

2012), both available in the BEAST2 software (Bouckaert et al., 2014). A reduced version of my sampling was used, the SNP matrices that had one individual per locality were used for the species delimitation analysis, again in order to reduce the time and computational requirements for this analysis. The SNP matrices used for this analysis had 50% missing data and they were filtered to only include bi-allelic SNPs (1000 maximum) (Ruby script available at https://raw.githubusercontent.com/mmatschiner/snapp_prep/master/snapp_prep.rb). This species delimitation method estimates a species tree and evaluates the species delimitation model all together, it also allows the comparison between models with different numbers of species and different assignments of samples to species. The model parameters were set to 1 for mutation rates u and v, and they were not sampled, the coalescent rate was set to 10 and it was sampled to consider the population size. Moreover, the likelihood correction was used to estimate the Bayes

33

factor (explained below). I analyzed the SNP data using the population size parameter, q (gamma

(1,250)), which results in a prior mean = 0.004, and l using a gamma distribution (2, 200) to accommodate the uncertainty in the parameter. To estimate the marginal likelihood (MLE), I used a path sampling and stepping-stone with 48 steps, the length of MCMC was 1,000,000, and number of samples discarded before the first recorded step set to 100,000 (pre-burnin). The models were ranked according the MLE and the Bayes factor (BF) is used for the model selection, calculated using the difference between the MLE of two models multiplied by two (BF

= 2 x (MLE of model 1 – MLE of model 2)). Positive BF values indicate support of model 1 and negative BF favor model 2. The BF scale is: 0

610 decisive (Leaché et al., 2014).

Maps

To understand the distribution of the Chiasmocleis albopunctata species group and to help define the species delimitation models, the distribution maps were made with R (R

Development Core Team, 2017) using the packages raster (Hijmans & Etten, 2012), rgdal

(Bivand et al., 2015a), maptools (Bivand et al., 2015b), and RColorBrewer (Neuwirth, 2014), and QGIS software (QGIS Development Team, 2015).

Results

34

3RADseq processing

My genetic sampling included 197 tissue samples from 54 localities of Chiasmocleis albopunctata (Table 2). After demultiplexing and filtering low quality reads, I obtained an average of ~1.5 million reads per sample and 111,204 of total prefiltered loci. My iPyRAD pipeline generated 51,451 total filtered loci and 10,605 loci on average per sample. After the alignment of the loci across samples, the final data matrices had between 8,137 and 2,861 loci depending on the amount missing data allowed (Table 6).

Phylogeography

The phylogenetic tree topology generated in RAxML for my SNP dataset using different amounts of missing data were similar, the major difference was the value of branch support. The

SNP matrix with 20% of missing data (21,173 SNPs) had better support in the major clades when compared to the 50% missing data (53,511 SNPs). Also, another discrepancy between them, was one of the clades (Chiasmocleis mehelyi) had a different placement in the maximum likelihood tree, which was the oldest divergent clade in the species tree and it was recovered in the coalescent tree (results below). According to my results, the C. albopunctata clade is comprised of five species: C. albopunctata, C. mehelyi, C. centralis, C. bicegoi, and C. sp one undescribed species (Fig. 6), I could compare my samples and determine the species clade with the recently published study de Sá et al. (2019), because some were the same samples. The

STRUCTURE results (unlinked loci with 50% of missing data allowed) for C. albopuntata recovered K=6 population clusters (Fig. 7), the DK statistic peaked for four and six populations clusters (Evanno et al., 2005). C. centralis have two clades and two genetic clusters (Fig. 6 and

7). The species tree generated using coalescent methods in SVDquartets (unlinked loci with 50%

35

and 20% of missing data) recovered the same geographic and genetic clades for the five species with high support (Fig. 8 and 9). In both trees the placement of C. mehelyi was the oldest divergent clade in the tree and the new species being a sister species of C. bicegoi (Fig. 8 and 9).

The distribution of all species shows some biogeographic pattern, C. centralis has a wide and central distribution on the Neotropical savannas being endemic of the Cerrado, and the other species are distributed in the periphery of the Cerrado (Fig. 10).

Divergence time

The species tree obtained from SNAPP had some differences when compared to the trees obtained from SVDquartets, also a coalescent method (Fig. 11). The tree recovered a similar topology proposed in previous studies except for the presence of a new species and it had a similar topology when compared to RAxML species tree: one clade with C. albopunctata being sister of C. mehelyi, C. mehelyi is the sister species of C. sp, C. bicegoi and C. centralis (Fig. 11).

The diversification of the Cerrado and Chaco species probably occurred in the Miocene around

~14 (million years ago (Mya)). The divergence time of C. albopuntata was ~14 Mya (95% HPD:

10.53 – 18.85 Mya), the time of divergence C. mehelyi was ~6.70 Mya (95% HPD: 2.29 – 12.18

Mya) (Fig. 11). The divergence time C. sp. was ~2.88 Mya (95% HPD: 1.089 – 5.16 Mya), and for C. bicegoi and C. centralis it was ~ 1.03 Mya (95% HPD: 0.32 – 2.02 Mya) (Fig. 11).

Species delimitation

Observing the results of the STRUCTURE analysis and RAxML tree, I ran seven different models lumping and splitting species: (A) current and the potential new species (C. sp); (B) lump of Chiasmocleis albupunctata and C. mehelyi; (C) lump of C. albupunctata and C.

36

mehelyi, and another lump of C. sp and C. bicegoi; (D) lump of C. sp and C. bicegoi; (E) lump of

C. sp, C. bicegoi and C. centralis; (F) split of C. centralis; (G) lump of C. sp and C. bicegoi, and split of C. centralis. The current taxonomy model with five species was rejected in favor of the three species model that lumps C. sp, C. bicegoi and C. centralis into one species, with strong support (Table 4 and Fig. 12).

Table 4: Results for BFD* species delimitation model selection. Model Species MLE Rank BF A: current taxonomy + Csp 5 -500.19 7 -- B: lump CA and CM 4 -114.41 3 -771.56 C: lump CA and CM, lump Csp and CB 3 -478.73 5 -42.93 D: lump Csp and CB 4 -69.30 2 -861.79 E: lump CC, Csp and CB 3 17.99 1 -1036.38 F: split CC 6 -463.85 4 -72.68 G: split CC and lump Csp and CB 5 -483.55 6 -33.28 CA= C. albopunctata, CM= C. mehelyi, CB= C. bicegoi, CC= C. centralis, Csp= C. sp.

Discussion

The Chiasmocleis albopunctata group showed strong biogeographic structure with six well-supported clades and six genetic clusters. The phylogenetic tree recovered is similar to that described in a previously published study (de Sá et al., 2019), with the exception of the potential presence of one new species. The distributions of the species C. albopunctata, C. mehelyi, C. bicegoi, and C. centralis were very poorly understood, and my sampling covered most of the distribution of the Cerrado and Chaco’s Chiasmocleis species (Forlani et al., 2011), including samples from the Bolivian savannas. The results presented here conclusively to address the distribution (Wallacean) knowledge gap. My results corroborate that Chiasmocleis centralis is endemic to the Cerrado (Valdujo et al., 2012), having a wide and central distribution in the

Neotropical savannas, comprising two clades and two genetic clusters inhabiting the Central

37

region (plateau) and the northern area of the Cerrado. In addition, the other Chiasmocleis species are geographically restricted to the southern periphery of the biome, supporting the hypothesis of diversification by allopatric speciation suggested for the genus (de Sá et al., 2019), and suggests a history of isolation and reinvasion of southern plateau areas during the climatically favorable periods (Gallardo, 1979). Such speciation by vicariant events is known to have shaped the diversity of squamates in the Neotropical savannas, resulting high levels of endemicity

(Nogueira et al., 2011). Recent studies have shown the effects of vicariant events on the genetic variation of the lizards Dermatonotus muelleri and Phyllopezus pollicaris for which the Central

Brazilian Plateau constituted a geographic barrier and genetically structured populations in the southwest and northeast of Brazil (Werneck et al., 2012a; Oliveira et al., 2018). Thus, knowing now the distribution of members of the C. albopunctata species group and phylogenetic relationships among them, it is possible to tease apart the phylogeography of the group and explore diversification patterns.

The divergence of Chiasmocleis albopuntata and C. mehelyi occurred during the

Miocene, followed by the recent divergence of Chiasmocleis sp., C. bicegoi and C. centralis during the Pliocene and Pleistocene. My results corroborate the historical biogeography of C. albopuntata species group described by de Sá et al. (2019), which connects old lineages from

Amazon Forest to Atlantic Forest, with subsequent connections between Cerrado and Chaco. The

C. albopunctata species group originated from the southern Atlantic Forest clade (de Sá et al.,

2019). The species C. albopunctata itself is distributed in the western region of the Cerrado, including the Bolivian savannas and Chaco, and it is the sister species of the other Cerrado and

Chaco species (i.e. C. mehelyi, C. sp, C. bicegoi and C. centralis). Haffer (1979) suggested a connection between Amazon and Atlantic Forest throughout the Cerrado, however, my results

38

corroborates with evidences of dispersal from the Atlantic Forest to Neotropical savannas

(Brazilian and Bolivian) and Chaco. Likewise, my results could support the connections between the Atlantic Forest and Amazon through a south-eastern forest corridor (Ledo & Colli, 2017), which might explain the diversification of the species in the south-eastern periphery of the biome. Additionally, during the middle Miocene (~13 Mya) the presence of the Paranaense sea in southern South America (Gross et al., 2016) was important for the diversification of Chacoan amphibian species and may represent one of the earliest factors influencing the diversification of the C. albopunctata species group. According Gallardo (1979), amphibian species extended its distribution from east to the west, i.e. dispersal to Chaco, and adapted to arid climatic conditions.

Thus, Chiasmocleis species may have extended its distribution from southern Atlantic Forest to

Chaco. Connections between the Amazon and Atlantic Forests through the Cerrado and Chaco have also been inferred for other vertebrate species for which the center of origin was the

Amazon Forest, followed by dispersal to the Atlantic Forest, with subsequent adaptation to dry open vegetation (Chaco and Neotropical savannas) (Costa, 2003; Lynch Alfaro et al., 2012;

Fouquet et al., 2014; Gehara et al., 2014).

The divergence time of Chiasmocleis albopunctata (~ 14 Mya, 95% HPD: 10.53 – 18.85

Mya) coincides with a dramatic cooling phase of global temperatures in the Miocene between 16

– 14 Mya. Periods of drier, colder climates are associated with lowering of the sea level given rise to shallow inland lakes, producing favorable environments for dispersal of species inhabiting open habitats (Haffer, 1979). There evidences of the presence of marine transgression

(Paranaense) in the Chaco-Pampean area during the middle and late Miocene (~ 13 - 5 Mya)

(Cione et al., 2011). Moreover, during the Neogene the Brazilian and Guiana Shields experienced further upwarping (Haffer, 1974), the Andes continued their orogenic growth (23 –

39

7 Mya), and the marine transgression (20 – 5 Mya) divided South America in to two regions, north-western and south-western (Haffer, 1974; Morrone, 2014). This presents a complex landscape in which the diversification of Chiasmocleis albopunctata to the other savanna- inhabiting species took place. Notably, the divergence of C. sp. (~ 2.9 Mya, 95% HPD: 1.089 –

5.16 Mya) coincided with another cooling period during the Pliocene (5 – 2 Mya) (Pearson &

Palmer, 2000), and the lowering of sea level. During the Pliocene (3.3 Mya) the climate was drier and colder (Dolan et al., 2015), and around 3.2 - 3 Mya the climatic conditions became warmer than the current climatic conditions according Hill (2015), followed by colder and dryer periods until the Last Glacial Maximum (21 Kya). Thus, these climatic shifts in the Quaternary resulted in expansion and retractions of suitable habitats for C. sp, C. bicegoi and C. centralis (~

1.03 Mya (95% HPD: 0.32 – 2.02 Mya)), which could have generated appropriate conditions for speciation. By the end of the Miocene (~ 5 Mya) the Paranaense sea disappeared (Cione et al.,

2011). Also, during the late Miocene and early Pliocene around 5 Mya the further upwarping of the Brazilian shield promoted the subsidence of the peripheral depression (Chaco and Cerrado), and also the immediate surroundings of the Andes were below the sea level through the

Cenozoic and did not emerge until the late uplift of Andes around 5 - 2 Mya (Mendes & Petri,

1971; Haffer, 1974; Colli, 2005; Morrone, 2014), could have stimulated the emergence of species distributed in the Cerrado savannas (i.e. C. sp, C. bicegoi, and C. centralis). Thus, the climatic shifts in combination with the geological events could be responsible for diversification of the Chiasmocleis albopunctata species group.

The species delimitation model selected contradicts the current taxonomy and favors a model with fewer species. The selected model had three species against five species present in the current taxonomy (four species) plus Chiasmocleis sp. This result differs from what is

40

commonly observed in studies of widespread species groups in the Neotropical region, which is a very rich in biodiversity and harbors a large number of cryptic species (Colli et al., 2002; Klink

& Machado, 2005; Nogueira et al., 2011; Gamble et al., 2012; Valdujo et al., 2012; Domingos et al., 2014; 2017; Gehara et al., 2014; Guarnizo et al., 2016). However, these species delimitation analyses are very sensitive to the choice of prior for critical parameters, such as q, high prior means will favor models with few species and group populations together, because it characterizes a probability that the average divergence within populations is somewhat high

(Leaché et al., 2014). The use of coalescent methods for genome-wide data is robust (Fujita et al., 2012; Leaché et al., 2014), however it requires some pre-processing of the data to estimate the priors that can improve the analysis. Furthermore, the use of a reduced dataset can bias the outcome of the analysis, producing simplistic phylogenetic relationships between the species or populations. The current research using genetic and morphologic data presented C. bicegoi and

C. centralis as two distinct species, it could be due the wide altitudinal and latitudinal distribution which results in a high phenotypic plasticity. Also, the use of few genetic markers could be less accurate than genome wide data, which could not recover C. bicegoi and C. centralis as one species.

Conclusion

This exploration of phylogenetic and phylogeographic patterns in the Chiasmocleis albopunctata species group addressed a number of deficiencies in our knowledge of Cerrado diversity and the factors influencing this taxon’s evolutionary history. The species group is widely distributed in the Bolivian and Brazilian savannas and Chaco, and there are at least tree species C. albopunctata, C. mehelyi, and a lump of C. bicegoi, C. centralis, and C. sp. The

41

results presented here highlight the likely influence of the Atlantic Forest as the source for species inhabiting the dry open areas of the Neotropical savannas.

Future work with this group will focus on diversifying the species delimitation analytical methods. Furthermore, the use of species distribution modeling could improve our understanding the species distribution changes in response to climatic shifts, allowing more nuanced analysis of demographic history.

42

FIGURES

Figure 6. Maximum likelihood tree for Chiasmocleis albopunctata species group obtained from RAxML (20% missing data). The localities are represented by the Code (Table 2) and species are represented by the colors: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange = C. centralis.

43

Figure 7. A) Population assignment results from processing of STRUCTURE analyses using the Evano DK method (Evanno et al., 2005) to identify the number of distinct genetic groups within the sample of Chiasmocleis; B) STRUCTURE bar plot illustrating probability of assignment to the six genetic clusters found for the Chiasmocleis albopunctata species group. Species are represented by the colors: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange and brown = C. centralis.

44

Figure 8. Coalescent species tree of the Chiasmocleis albopunctata species group obtained from SVDquartets.

45

Figure 9. Coalescent species tree of the Chiasmocleis albopunctata species group obtained from SVDquartets (Table 2). Species are represented by the colors: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange = C. centralis.

46

Figure 10. Distribution map of the member of the Chiasmocleis albopunctata species group in which sampling localities for each species are represented by the colored circles: green = C. albopunctata; red = C. mehelyi; yellow = C. sp; blue = C. bicegoi, and orange = C. centralis; red outline delimits the boundaries of the Cerrado biome; numbers correspond to localities listed in Table 2; the gray scale is elevation (m).

47

Figure 11. Coalescent model species tree for the Chiasmocleis albopunctata species group with divergence time estimates obtained from SNAPP; horizontal gray bars represent the 95% credibility interval of divergence time estimates at each node; numbers above branches indicate posterior probability for monophyly of the descendant clade. The x-axis represents the time before present (Mya).

48

Figure 12. Species tree obtained from species delimitation analysis for Chiasmocleis albopunctata species complex and the three species model that lumps C. centralis, C. bicegoi and C. sp into one species. Posterior probability is shown on the branch.

49

CHAPTER 2

COMPARATIVE PHYLOGEOGRAPHY OF THREE BRAZILIAN SAVANNA FROGS

INTRODUCTION

The biodiversity observed today in the Neotropical region is the result of a complex evolutionary past on a dynamic landscape that has been influenced by dramatic shifts in habitats, climate, and a complex geology, which resulted in a diverse and rich herpetofauna (Duellman,

1979; Avise et al., 2016). Phylogeography seeks to understand inferred patterns of evolutionary relationship with respect to the geographic distribution of the samples/populations in a phylogeny (Avise et al., 1987). This field fills an important gap between the study of deeper evolutionary events (historical biogeography) and fine-scale, ongoing demographic processes

(population genetics) (Bermingham & Moritz, 1998). Most of the phylogeographic studies in the

Neotropical region have focused on the Amazon and Atlantic Forest because these regions are known for their high levels of biodiversity. On the other hand, the Cerrado savannas were considered to be poor in biodiversity until three decades ago (Heyer, 1988; Vitt, 1991), and through the intensification of studies focused on the Neotropical savannas, it has been revealed that this region is very diverse with high levels of endemicity (Colli et al., 2002; Klink &

50

Machado, 2005; Nogueira et al., 2011; Valdujo et al., 2012). While these regions are now known to be incredibly diverse and home to a unique biota, the patterns of diversification and the historical factors influencing them are still poorly known. Lately, phylogeographic studies have been conducted in the Cerrado, highlighting unique evolutionary and diversification patterns that have shaped the region's biodiversity we observe today.

Recent studies of amphibians in the Neotropical savannas have demonstrated some concordant phylogeographic patterns. The geomorphological changes of the Miocene and

Pliocene, and the climatic fluctuations of the Pleistocene have been implicated as important events in the diversification of a number of species (Santos et al., 2009; Maciel et al., 2010;

Vallinoto et al., 2010; Prado et al., 2012; Gehara et al., 2014). During the Miocene – Pliocene it is clear that the uplift of the Andes and the formation of the Amazonian basin, as well as the dispersal of open/dry habitat adapted species from the Guiana Shield, Andes, and Brazilian

Shield were responsible for the assembly of much of the beta diversity of the region (Santos et al., 2009; Maciel et al., 2010; Gehara et al., 2014). With the relatively ancient uplift of the

Brazilian Shield in the late Miocene (23 – 5 Mya) dispersal routes and barriers were greatly altered, changing South American communities on a continental scale. The effects of subsequent climatic fluctuations in the Pleistocene caused the expansion and contraction of biomes, such as the Cerrado (Werneck et al., 2012b). However, it seems that the Cerrado was largely limited to the Brazilian Central Shield, meaning these cycles of expansion and contraction may have promoted the diversification in situ in the Brazilian Central Shield (Maciel et al., 2010; Prado et al., 2012). As an example, the connection between the western Amazon Forest and the south- eastern Atlantic Forest during the Pleistocene, currently separated from one another by the

51

Cerrado (Fig. 2), would have bisected this extensive savanna, an important event for the diversification of species adapted to the savanna’s dry and open habitats (Ledo & Colli, 2017).

A close relationship between species restricted to the Cerrado and their relatives in the adjacent Atlantic Forests has been frequently supported in biogeographic studies. The most recent common ancestor of the Cerrado and Chaco Chiasmocleis species arose from the southern

Atlantic Forest clade during the Miocene (19 Mya) (de Sá et al., 2019). A similar pattern was reported for frogs of the genus Adenomera (Fouquet et al., 2014), for which both the Atlantic

Forest clade and the Dry Diagonal (Cerrado, Chaco, and Caatinga) clade emerged at the same time, and the common ancestor of the Atlantic Forest clade could be related to an ancient

Amazon and Atlantic Forest. Gehara et al. (2014), also recovered a similar pattern for frogs of the Dendropsophus minutus species group, for which the center of origin was inferred to be the

Amazon region with dispersal to eastern Brazil (Atlantic Forest), followed by expansion from the east of Brazil into Bolivia. Each of these studies have focused on anuran amphibians, whose biphasic life cycle makes them well-suited to exploring the influence of changes in the extent of dry habitats throughout time. Remarkably, the congruent patterns have been reported from independent studies with various amphibian species that do not explicitly implement a comparative phylogeographic approach, highlighting an opportunity for comparative studies of anuran diversification in the region.

Comparing the phylogeography of different taxa, especially those that vary in their life history strategies, can reveal fundamental associations between population processes and regional patterns of diversity and biogeography (Bermingham & Moritz, 1998). Therefore, comparative phylogeographic analysis emerged because researchers were looking for congruence in phylogeographic patterns to test and apply general evolutionary

52

processes/hypotheses for a particular region. In the Neotropical region, again the Amazon and

Atlantic Forests were the focus of attention due the high levels of biodiversity, hence the number of comparative phylogeographic studies are numerous, especially when compared to the Cerrado

(e.g. Cerrado: Collevatti et al., 2011; Françoso et al., 2016)(e.g. Altantic Forest: De Moraes-

Barros et al., 2006; Carnaval et al., 2009; Prates et al., 2016b; a)(e.g. Amazon: Costa, 2003;

Solomon et al., 2008). However, none of the studies in the Cerrado were conducted using amphibian species. In the few comparative studies of the Neotropical savannas, the evolutionary patterns observed are primarily associated with the climatic fluctuation during the Quaternary

(Collevatti et al., 2011; Françoso et al., 2016), with evidence of contractions in the last interglacial (LIG) and expansion during the last glacial maximum (LGM) of the species distribution. In addition, with molecular data evidence of bottleneck followed by rapid expansion

(Françoso et al., 2016). Nonetheless, it is observed incongruence in the topology of the phylogenetic trees due differences in species-specific traits was observed (Françoso et al., 2016).

Recent advances in the resolution with which genomic-scale phylogeographic data and information about the species-specific environmental tolerances could improve comparative phylogeographic studies, and contribute to our understanding of concordance and discordance among patterns of diversification (Zamudio et al., 2016).

Objective

In order to explore the factors influencing diversification and population structure in the

Cerrado under a comparative framework, I investigated the comparative phylogeography of

Chiasmocleis centralis, Dendropsophus rubicundulus and Physalaemus nattereri. These species

53

belong to different families and have very different morphology, and life-history strategies.

Therefore, producing a comparative phylogeographic study with these taxa that exhibit similar distributions in the Cerrado savanna allowed me to elucidate the general evolutionary patterns of this unique region. Observing the lack of comparative phylogeographic studies in the

Neotropical savannas and the differences between the chosen species, the aim of this chapter is to compare the phylogeography of Chiasmocleis centralis, Dendropsophus rubicundulus and

Physalaemus nattereri. Congruence of phylogeographic patterns across such different anuran taxa would provide strong evidence in support of the common responses of lineages within the region to the environmental factors invoked by hypotheses proposed to explain evolutionary patterns.

Hypothesis to be tested

• The uplift of the Brazilian plateaus and landscape fragmentation during the Neogene, and

climatic fluctuations during the Quaternary were important events for the diversification

and the distribution of the genetic diversity in populations of Chiasmocleis centralis,

Dendropsophus rubicundulus, and Physalaemus nattereri.

Predictions

a. Knowing that Chiasmocleis centralis, Dendropsophus rubicundulus, and

Physalaemus nattereri are mainly distributed in the Cerrado, it is expected that

intraspecific patterns of divergence will be geographically coincident (e.g. shared

breaks in genetic continuity across the landscape) as these species would have

responded similarly to recent (Pleistocene) changes in habitat continuity.

54

b. Facing the geomorphological changes in central Brazilian shield, i.e. uplift of

plateau and subsidence of peripheral depressions, it is expected high genealogical

structure and genetic diversity in the plateaus and less genetic diversity is

expected in the valleys (Plateau vs. valley hypothesis (Werneck, 2011)).

c. Given the climatic fluctuations during the Quaternary and the formation of refugia

in the Cerrado biome, high genetic diversity is expected in stable areas (refugial),

and less genetic diversity is expected in unstable areas (recent re-colonization,

low genetic diversity).

Methods

Sampling

During my fieldwork in December of 2016, in Brasília, Federal District, Brazil, I actively searched for and manually captured individuals of the Chiasmocleis centralis, Dendropsophus rubicundulus, and Physalaemus nattereri. Collected individuals were euthanized with lidocaine ointment or lethal injection of xylocaine. Tissue samples of muscle and/or liver were collected and preserved in cryogenic tubes with 95% ethanol until they could be placed in a -80º C freezer.

Specimens were then fixed in 10% formalin and preserved in 70% ethanol. All specimens and tissues collected were deposited in the Coleção Herpetológica da Universidade de Brasília

(CHUNB), Brasília – DF, Brazil. Also, I gathered tissue samples of C. centralis, D. rubicundulus, and P. nattereri from herpetological collections in Brazil, Paraguay, Argentina and

Germany.

55

For the study of phylogeographic patterns of each of the focal species I gathered 91 tissue samples from 25 localities of C. centralis (species of C. albopunctata group) (Table 5), 115 tissue samples from 39 localities of Dendropsophus rubicundulus (Table 6), and 181 tissue samples from 52 localities of Physalaemus nattereri (Table 7). Furthermore, for phylogenetic analyses I used two samples of Hypsiboas albopunctatus, D. marmoratus, D. leucophyllatus, D. minutus, and D. jimi as outgroups of D. rubicundulus; Ctenophryne geayi, Dermatonotus muelleri, and Elachistocleis ovalis as outgroups of C. centralis; Pleurodema diplolister, P. centralis, P. cuvieri, and Pseudopaludicula mystacalis as outgroups of P. nattereri.

56

Table 5: Population localities of Chiasmocleis centralis samples used in this study. The number (#) corresponds to the localities in the map (Fig. 21). The Code is used as the locality code and the colors represent the population cluster in the phylogenetic trees presented in the results (Fig. 14). # Code Municipality Country State N ind Longitude Latitude Cluster 1 chapa Chapada dos Guimarães Brazil MT 1 -55.5399 -15.1084 West 2 tesou Tesouro Brazil MT 4 -53.4580 -15.8890 West 3 ribei Ribeirão Cascalheira Brazil MT 4 -52.0810 -12.8579 West 4 brasi Brasília Brazil DF 8 -47.8761 -15.9396 Central-East 5 luzia Luziânia Brazil GO 5 -48.0251 -16.6749 Central-East 6 graom Grão Mogol Brazil MG 1 -42.5781 -16.7372 Central-East 7 sanri Santana do Riacho Brazil MG 3 -43.6039 -19.3435 Central-East 8 curve Curvelo Brazil MG 8 -44.4487 -18.8207 Central-East 9 burit Buritizeiro Brazil MG 3 -44.9619 -17.3508 Central-East 10 sanfe Santa Fé Brazil MG 1 -45.4139 -16.6900 Central-East 11 bonit Bonito de Minas Brazil MG 5 -44.7500 -15.2500 Central-East 12 peixe Peixe Brazil TO 6 -48.5552 -11.9968 Central-North 13 paran Paranã Brazil TO 6 -47.8122 -12.6959 Central-North 14 sdomi São Domingos Brazil GO 1 -46.4870 -13.5370 Central-North 15 altop Alto Paraíso de Goiás Brazil GO 8 -47.5352 -14.1830 Central-North 16 nique Niquelândia Brazil GO 1 -48.4018 -14.4985 Central-North 17 palma Palmas Brazil TO 3 -48.1536 -10.2186 Central-North 18 lajea Lajeado Brazil TO 4 -48.2899 -9.8483 Central-North 19 casea Caseara Brazil TO 2 -49.9587 -9.3044 Central-North 20 goiat Goiatins Brazil TO 3 -47.4811 -8.0936 Central-North 21 palte Palmeirante Brazil TO 1 -48.1402 -7.8898 Central-North 22 carol Carolina Brazil MA 8 -47.4136 -7.2980 Central-North 23 estre Estreito Brazil MA 3 -47.2013 -6.7826 Central-North 24 babac Babaculândia Brazil TO 1 -47.8559 -7.1751 Central-North 25 palmt Palmeiras do Tocantins Brazil TO 1 -47.6425 -6.5997 Central-North N ind = number of individuals. MT= Mato Grosso; MG = Minas Gerais; GO = Goiás; DF= Distrito Federal; TO = Tocantins; MA = Maranhão.

57

Table 6: Population localities of Dendropsophus rubicunculus samples used in this study. The number (#) corresponds to the localities in the map (Fig. 22). The Code is used as the locality code and the colors represent the population cluster in the phylogenetic trees presented in the results (Fig. 15). # Code Municipality Country State N ind Longitude Latitude Cluster 1 santc Santa Cruz Bolivia - 2 -62.0001 -16.3596 South 2 lucas Lucas do Rio Verde Brazil MT 9 -56.1657 -13.0377 South 3 catal Catalão Brazil GO 1 -47.7282 -17.9231 South 4 sucup Sucupira Brazil TO 1 -48.8352 -12.1133 South 5 doisi Dois Irmãos do Tocantins Brazil TO 1 -49.0665 -9.2714 South 6 goian Goianorte Brazil TO 1 -49.0099 -8.8737 South 7 ribei Ribeirão Cascalheira Brazil MT 4 -52.1067 -12.8530 South 8 brita Britânia Brazil GO 3 -51.0905 -15.0990 South 9 caiap Caiapônia Brazil GO 1 -51.8156 -17.0075 South 10 jarag Jaraguá Brazil GO 3 -49.3854 -15.7052 Central 11 piren Pirenópolis Brazil GO 2 -49.0361 -15.7573 Central 12 brasi Brasília Brazil DF 5 -47.6983 -15.5892 Central 13 forsa Formosa Brazil GO 1 -47.4246 -15.3245 Central 14 sjoao São João d’Aliança Brazil GO 12 -47.4230 -14.4857 Central 15 flore Flores de Goiás Brazil GO 4 -46.9046 -14.6517 Central 16 alvor Alvorada do Norte Brazil GO 2 -46.6315 -14.5665 Central 17 caval Cavalcante Brazil GO 6 -47.4322 -13.8115 Central 18 sdomi São Domingos Brazil GO 3 -46.4870 -13.5370 Central 19 minac Minaçú Brazil GO 1 -48.3488 -13.4988 Central 20 paran Paraña Brazil TO 4 -47.8122 -12.6959 Central 21 arrai Arraias Brazil TO 3 -47.0479 -12.8187 Central 22 sdesi São Desidério Brazil BA 4 -45.0872 -12.7056 Central 23 jabor Jaborandi Brazil BA 1 -45.1447 -13.9720 Central 24 burit Buritizeiro Brazil MG 3 -44.9619 -17.3508 Central 25 corin Corinto Brazil MG 1 -44.4346 -18.3585 Central 26 conmd Conceição do Mato Dentro Brazil MG 1 -43.6159 -19.0340 Central 27 sanri Santana do Riacho Brazil MG 7 -43.6717 -19.1281 Central 28 lagoa Lagoa Santa Brazil MG 3 -43.8871 -19.6280 Central 29 divin Divinópolis Brazil MG 1 -44.9311 -20.1242 Central 30 pirap Pirapetinga Brazil MG 1 -42.3640 -21.6794 Central 31 buri Buri Brazil SP 3 -48.5754 -23.7518 South 32 rioco Rio da Conceição Brazil TO 4 -52.7438 -10.9632 North 33 barre Barreiras Brazil BA 3 -45.0631 -12.1180 North 34 matei Mateiros Brazil TO 4 -46.7299 -10.5676 North 35 carol Carolina Brazil MA 2 -47.4215 -7.3385 North 36 loret Loreto Brazil MA 2 -45.0867 -7.0372 North 37 gonc Ribeiro Gonçalves Brazil PI 1 -44.6918 -7.4342 North 38 brejo Brejo do Piauí Brazil PI 1 -42.8046 -8.2053 North 39 pirip Piripiri Brazil PI 4 -41.7083 -4.1014 North N ind = number of individuals. MT= Mato Grosso; SP= São Paulo; MG = Minas Gerais; GO = Goiás; DF= Distrito Federal; TO = Tocantins; MA = Maranhão; BA = Bahia; PI = Piauí

58

Table 7: Population localities of Physalaemus nattereri samples used in this study. The number (#) corresponds to the localities in the map (Fig. 23). The Code is used as the locality code and the colors represent the population cluster in the phylogenetic trees presented in the results (Fig. 16). # Code Municipality Country State N ind Longitude Latitude Cluster 1 santc Santa Cruz Bolivia - 2 -62.0001 -16.3596 West 2 vilab Vila Bela da Santíssima Trindade Brazil MT 11 -60.0491 -14.8853 West 3 vales Vale de São Domingos Brazil MT 3 -58.9520 -14.9749 West 4 arapu Araputanga Brazil MT 2 -58.4575 -15.2563 West 5 tanga Tangará da Serra Brazil MT 9 -58.3230 -14.4314 West 6 sapez Sapezal Brazil MT 2 -58.6535 -13.1522 West 7 juina Juína Brazil MT 6 -58.1238 -12.3354 West 8 lamba Lambari D'Oeste Brazil MT 5 -57.7469 -15.1599 West 9 cacer Caceres Brazil MT 3 -57.8370 -16.5285 West 10 sjose São José do Rio Claro Brazil MT 7 -56.7214 -13.4467 West 11 rosar Rosário D'Oeste Brazil MT 1 -56.6335 -15.1199 West 12 varze Várzea Grande Brazil MT 1 -56.2550 -15.5709 West 13 nossa Nossa Senhora do Livramento Brazil MT 1 -56.3747 -16.2399 West 14 salev Santo Antônio do Leverger Brazil MT 1 -55.4397 -16.4568 West 15 prima Primavera do Leste Brazil MT 3 -54.1746 -15.3332 West 16 gauch Gaúcha do Norte Brazil MT 1 -53.0872 -13.2475 West 17 ribei Ribeirão Cascalheira Brazil MT 4 -52.1067 -12.8530 North 18 altoa Alto Araguaia Brazil MT 3 -53.4692 -17.4450 Central 19 guira Guiratinga Brazil MT 2 -53.7583 -16.3494 Central 20 tesou Tesouro Brazil MT 4 -53.2164 -15.8074 Central 21 campo Campo Grande Brazil MS 2 -54.7259 -20.5156 Central 22 teren Terenos Brazil MS 3 -54.8603 -20.4422 Central 23 conce Concépcion Paraguay - 5 -57.3540 -22.5871 Central 24 amamb Amambay Paraguay - 10 -56.2581 -22.6719 Central 25 canin Canindeyú Paraguay - 2 -54.3333 -24.0167 Central 26 cosmo Cosmopolis Brazil SP 5 -47.1861 -22.6515 Central 27 luiza Luiz Antônio Brazil SP 2 -47.8480 -21.5830 Central 28 pedre Pedregulho Brazil SP 1 -47.4361 -22.2247 Central 29 icem Icem Brazil SP 1 -49.1859 -20.3664 Central 30 sacra Sacramento Brazil MG 2 -47.3576 -19.8869 Central 31 brasi Brasília Brazil DF 9 -47.6848 -15.6414 Central 32 luzia Luziânia Brazil GO 5 -47.7141 -16.2613 Central 33 piren Pirenópolis Brazil GO 1 -49.0136 -15.8110 Central 34 petro Petrolina de Goiás Brazil GO 2 -49.3139 -16.1119 Central 35 burit Buritizeiro Brazil MG 4 -44.9619 -17.3508 North 36 sanfe Santa Fé Brazil MG 2 -45.4139 -16.6900 North 37 flore Flores de Goiás Brazil GO 5 -46.9046 -14.6517 North 38 alvor Alvorada do Norte Brazil GO 4 -46.6315 -14.5665 North 39 caval Cavalcante Brazil GO 1 -47.4322 -13.8115 North 40 sdomi São Domingos Brazil GO 3 -46.4870 -13.5370 North 41 sdesi São Desidério Brazil BA 2 -44.6135 -13.0173 North 42 paran Paraña Brazil TO 8 -47.8122 -12.6959 North 43 rioco Rio da Conceição Brazil TO 3 -46.7491 -11.3438 North 44 peixe Peixe Brazil TO 7 -48.5552 -11.9968 North 45 figue Figueirópolis Brazil TO 3 -48.9616 -12.1843 North 46 gurup Gurupi Brazil TO 1 -48.8867 -11.6521 North 47 sucup Sucupira Brazil TO 1 -48.8352 -12.1133 North 48 lajea Lajeado Brazil TO 5 -48.2899 -9.8483 North 49 palte Palmeirante Brazil TO 1 -48.1402 -7.8898 North 50 carol Carolina Brazil MA 7 -47.1426 -7.2433 North 51 estre Estreito Brazil MA 2 -47.2013 -6.7826 North 52 palmt Palmeiras do Tocantins Brazil TO 1 -47.6425 -6.5997 North N ind = number of individuals. MT= Mato Grosso; MS = Matogrosso do Sul; SP= São Paulo; MG = Minas Gerais; GO = Goiás; DF= Distrito Federal; TO = Tocantins; MA = Maranhão; BA = Bahia.

59

Molecular methods

Whole genomic DNA was extracted from tissue samples (muscle and/or liver) using the

DNA salt extraction protocol (Aljanabi & Martinez, 1997). In order to explore phylogeographic patterns I used the triple digest, restriction-site-associated DNA sequencing (3RADseq) method to generate sequence data from thousands of loci for all individuals (Glenn et al., 2017).

Advances in high-throughput sequencing have proven highly effective in a wide range of studies such as high-resolution genetic mapping (Baird et al., 2008) and population genetics (Davey et al., 2010), but has proven particularly effective in studies of phylogeography (Emerson et al.,

2010). This method has some advantages compared to previous methods: it does not require the prior development of any genomic resources (Emerson et al., 2010), it is an inexpensive and rapid method (Peterson et al., 2012), and it can be used for simple and complex analysis such as genotyping and single-nucleotide polymorphism (SNP) discovery, applicable to quantitative genetic and phylogeographic studies, respectively (Davey et al., 2010).

The 3RADseq method enables the identification, verification and scoring of thousands of

SNPs from anonymous regions in seven steps (Glenn et al., 2017):

1. Digestion – digestion of 100 ng of genomic DNA for 1 hour at 37º C with two restriction

endonucleases enzymes (XbaI and EcoRI) and a third enzyme (NheI), that digests the

dimers formed by the phosphorylation on the 5’ end of the iTru Read 1 adapter. These

enzymes were chosen based on an in silico digest of the genome of Xenopus laevis that

suggested the combination would yield an appropriate number of loci for downstream

analyses (i.e. not fewer than 2,000 loci, not more than 10,000 loci).

2. Ligation – DNA ligation with double-stranded adapters whose 'sticky' ends were

complimentary to those left by the restriction enzyme. The ligated adapters contain a

60

terminal region that is complimentary to primers that are used in step 4 below. Ligation

reactions were run using the following cycle: 22ºC for 20 min, 37ºC for 10 min, repeat

22°C and 37°C steps once, and 80ºC for 20 min).

3. Clean-up –SeraMag speedbeads were used, in combination with a 96-well plate magnet,

to purify the ligation reaction of any non-DNA molecules before next step. Ligation

reactions were washed twice with Ethanol and eluted in IDTE buffer.

4. Polymerase chain reaction (PCR) – All DNA fragments possessing an XbaI site on one

end and an EcoRI site on the other end were amplified using a combination of primers

(complimentary to the ligated adapter sequence) were amplified using iTru5 and iTru7

primers. These primers each possessed a unique barcode sequence that would enable

assignment of all reads to an individual frog. Each DNA sample was amplified with a

unique combination of iTru5/7 primer barcodes, such that each sample could be

identified. After amplification, PCR products were pooled in groups of approximately 24

samples with 5 to 10 ng per pool. Pools were then cleaned with the MinElute® PCR

Purification Kit (QIAGEN) to purify the PCR products.

5. Fragment size selection – 5 to 10 ng of the cleaned PCR product pool containing the

DNA fragments of interest were size separated on a Pippin PrepTM (Sage science) using

1.5% agarose gel cassettes. From this, fragments ranging in size from 328-418 base pairs

were separated and retained for sequencing.

6. Quantification of samples – Purified, size-selected fragment pools were quantified using

qPCR with the KAPA SYBR Fast qPCR kit on an Applied BiosystemsTM QuantStudioTM

real-time PCR system.

61

7. Sequencing – Pools were sequenced on Illumina NextSeq® 500/550 High Output Kit v2

(75 cycles) flow cells, at the National Center for Natural Products Research at the

University of Mississippi.

The DNA sequences obtained after the Illumina sequencing were processed using the following bioinformatic pipeline on a High-performance Computing Cluster at the Mississippi

Center for Supercomputing Research (MCSR). The raw sequences were demultiplexed by the individual barcodes and converted from .bcl to .fastq file format using bcl2fastq Conversion

Software available from Illumina (https://support.illumina.com/ sequencing/sequencing_software/bcl2fastq-conversion-software.html). The iPyRAD (v0.7.28) pipeline, a newer version of PyRAD, was used to perform de novo assembly of RADseq datasets

(Eaton, 2014), identifying loci within individuals and then creating alignments of loci across all individuals sequenced. The clustering similarity threshold was set to 0.85, and the offset for quality scores was left at the default value of 33, the minimum depth for statistical and the minimum depth majority rule for the consensus base calling was set to 10, and the maximum number of allowed barcode mismatches was set to 2 (our barcodes had, at minimum, 3 nucleotide degeneracy). For phylogenetic inference, I produced SNP datasets including the outgroups, and for population genetic inference, I produced SNP datasets without the outgroups.

To explore the effects of missing data, three different matrices were produced allowing 20%,

50%, and ~100% (i.e. no threshold was set) of missing data in the shared loci for all individuals

(Table 8).

62

Table 8: Output of iPyRAD for number of total filtered loci and SNPs including and not including outgroups. Species N Missing Min. ind. Loci Unlinked Total Outgroups data SNPs SNPs 20% 84 3,895 3,777 18,407 Yes 105 50% 52 9,040 8,643 39,871 Yes ~100% 4 36,407 25,979 96,939 Yes C. centralis 20% 72 5,511 5,081 19,119 No 91 50% 45 10,171 9,280 34,032 No ~100% 4 35,528 24,346 78,952 No 20% 100 1,519 1,507 16,865 Yes 125 50% 62 6,000 5,969 63,183 Yes D. ~100% 4 68,251 55,155 329,980 Yes rubicundulus 20% 92 2,191 2,170 21,580 No 115 50% 57 6,868 6,821 66,422 No ~100% 4 67,848 54,445 316,440 No 20% 153 4,455 3,984 14,970 Yes 190 50% 95 7,552 6,674 25,098 Yes ~100% 4 35,534 24,711 87,848 Yes P. nattereri 20% 144 5,013 4,423 15,358 No 181 50% 90 7,841 6,863 24,204 No ~100% 4 31,564 20,690 66,050 No N= total samples; Min. ind. = minimal number of individuals per locus.

Phylogenetics, Phylogeography, and Population structure

I used two approaches to infer the phylogeographic structure from my SNP data. First, I used RAxML v8.2 which performs a maximum likelihood-based inference of phylogenetic relationships using SNP matrices with 50% missing data (Stamatakis, 2014). In addition, I used

Lewis (2001) acquisition bias correction, a conditional likelihood method for dealing with variable sites. To remove the non-binary sites, the matrices with all SNPs were pre-processed using the package Phrynomics in R (Leaché et al., 2015; R Development Core Team, 2017). I used the K80 model of nucleotide substitution with ASC_GTRCAT rate heterogeneity. Branch support for recovered phylogenetic groupings was estimated using an automatic bootstrap function (autoMRE) (Pattengale et al., 2010). For the species-tree inference using SNPs, I used

63

SVDquartets (Chifman & Kubatko, 2014) to apply a multi-species coalescent model implemented in PAUP* v4.0 (Swofford, 2002). I used unlinked SNPs matrices, as described above, and with 12 million random quartets with 100 bootstrap replicates, setting the outgroups as monophyletic groups.

To infer population genetic structure, I used unlinked SNP matrices with 20% of missing data without outgroups for each species. The analysis is a model-based clustering method performed in STRUCTURE v2.3 (Pritchard et al., 2000), using the package ParallelStructure in R

(Besnier & Glover, 2013). First, for each species I estimated the value of lambda to be used in the subsequent runs performing 10 runs with K=1. After that, I implemented 5 runs for each value of K ranging from 1 to 8 (40 runs) with 100,000 burn-in steps and a Markov chain Monte

Carlo (MCMC) of 400,000 steps. I then ran the results of STRUCTURE through the STRUCTURE

HARVESTER online software (Earl & vonHoldt, 2012), which explores how many distinct genetic groups there are within the data set and whether there is evidence for dispersal and/or gene flow.

I then estimated the optimal K value by implementing the method of Evanno (Evanno et al.,

2005) and verified changes in the Ln P(D) of each model. I used CLUMPAK server to produce the graphics to illustrate STRUCTURE results (Kopelman et al., 2015).

Divergence time and Demographic history

In order to study the demography of all populations, I used the python package momi2

(MOran Models for Inference) (Kamm et al., 2018), which computes the sample frequency spectrum (SFS) and estimates the demographic history, including time of divergence, population size, migration, and bottlenecks. I used reduced SNP matrices with one sample per locality, a maximum of 20% missing data per sample allowed, and excluding outgroups, to reduce the time

64

and computational requirements for this analysis. I used the results from STRUCTURE and

RAxML to assign each sample to a population and create momi2 models. I created six models:

(1) divergence with no migration not considering the effective population size; (2) divergence with no migration considering the estimated effective population size; (3) divergence with bidirectional migration between populations A and B, considering the estimated effective population size; (4) divergence with bidirectional migration between populations B and C, considering the estimated effective population size; (5) divergence with unidirectional migration between populations C and B, and B and A, considering the estimated effective population size;

(6) divergence with unidirectional migration between populations B and A, B and C, considering the estimated effective population size (Fig. 13). I considered a typical vertebrate mutation rate of µ = 1 X 10-8 and an average generation time of 1.5 years, in order to make the results biologically meaningful (Barley et al., 2018).The time of diversification, population size and migration between populations were estimated using momi2 and the models were optimized using maximum likelihood. In addition, I used the f3-statistic to assess the presence of introgression between populations (Patterson et al., 2012). After that, I used the Akaike

Information Criterion (AIC) with 100 bootstrap replicates to compare the fit of the population models to the data and identify the model with the best fit.

Population genetic diversity and isolation by distance (IBD)

The population genetic diversity, i.e. average expected heterozygosity (Hs), inbreeding coefficient Fis and pairwise fixation index Fst (Reynolds et al., 1983), were calculated for the population clusters found in the Structure analysis (explained before in the methods section).

These parameters were estimated in R (R Development Core Team, 2017) using the packages

65

adegnet (Jombart, 2008) and hierfstat (Goudet, 2005). Furthermore, I used R (R Development

Core Team, 2017) to test for evidence of isolation by distance (IBD) (Wright, 1943) among sample localities by performing a Mantel test using a matrix regression between the genetic (Fst) and Euclidean geographic distances matrix with 999 permutations. For this analysis, I used the packages adegnet (Jombart, 2008) and ade4 (Dray & Dufour, 2007).

Species distribution models (SDMs)

The species niche modeling was performed to evaluate the effects of climate on the demography of Chiasmocleis centralis, Dendropsophus rubicundulus, and Physalaemus nattereri. For this purpose, the climatic data before present (bp): Pleistocene: late-Holocene

(Fordham et al., 2017), Meghalayan (4.2-0.3 Kya) (Fordham et al., 2017), mid-Holocene

(Fordham et al., 2017), Northgrippian (8.326-4.2 Kya) (Fordham et al., 2017), early-Holocene

(Fordham et al., 2017), Greenlandian (11.7-8.326 Kya) (Fordham et al., 2017), late- Pleistocene

Younger Dryas Stadial (12.9-11.7 Kya) (Fordham et al., 2017), late- Pleistocene Bølling-Allerød

(14.7-12.9 Kya) (Fordham et al., 2017), late- Pleistocene Heinrich Stadial 1 (17.0-14.7

Kya)(Fordham et al., 2017), Last Glacial Maximum (ca. 21 ka) (Karger et al., 2017), Last

Interglacial (ca. 130 Kya) (Otto-bliesner et al., 2009), Marine Isotope Stage 19 (MIS19) in the

Pleistocene (ca. 787 Kya) (Brown et al., 2018); Pliocene: mid-Pliocene warm period (3.264-

3.025 Mya) (Hill, 2015), Marine Isotope Stage M2 (ca. 3.3 Mya) (Dolan et al., 2015); and current, (1979 – 2013): Anthropocene (Karger et al., 2017) were obtained at the Paleoclim website (http://paleoclim.org) (Brown et al., 2018), and the altitude was obtained from Worlclim website (http://worldclim.org) (Hijmans, 2005). All environmental data were obtained and analyzed in raster format, resolution 2.5’, in WGS84 horizontal datum. I collected 20

66

environmental variables: altitude, annual mean temperature, mean diurnal range (mean of monthly (maximum temp - minimum temp)), isothermality, temperature seasonality (standard deviation *100), maximum temperature of warmest month, minimum temperature of coldest month, temperature annual range, mean temperature of wettest quarter, mean temperature of driest quarter, mean temperature of warmest quarter, mean temperature of coldest quarter, annual precipitation, precipitation of wettest month, precipitation of driest month, precipitation seasonality (coefficient of variation), precipitation of wettest quarter, precipitation of driest quarter, precipitation of warmest quarter, and precipitation of coldest quarter.

I used the dataset of presence records of Chiasmocleis centralis, Dendropsophus rubicundulus, and Physalaemus nattereri, i.e the geographic coordinates of the samples used in this study (Table 5, 6 and 7). I performed a principal component analysis (PCA) to exclude any correlated environmental variables, using the program R (R Development Core Team, 2017).

Variables with similar values of PCA loadings in the first and second components were considered correlated and removed from further analysis, I retained one of the set of highly correlated variables. The included variables were: (1) altitude, (4) temperature seasonality

(standard deviation *100), (8) precipitation seasonality (coefficient of variation), (10) precipitation of driest quarter, (11) precipitation of warmest quarter, and (12) precipitation of coldest quarter, (15) precipitation seasonality (coefficient of variation), (16) precipitation of wettest quarter,(17) precipitation of driest quarter, (18) precipitation of warmest quarter, and (19) precipitation of coldest quarter.

I used the maximum entropy algorithm (Maxent) to build SDMs for Chiasmocleis centralis, Dendropsophus rubicundulus, and Physalaemus nattereri for all species separately.

Maxent is an algorithm that uses presence-only distribution records and environmental variables

67

to build a SDM (Phillips et al., 2006), representing areas of predicted habitat suitability (Peterson

& Vieglais, 2001). To validate the SDMs using bootstrap, I randomly split the occurrence points into training data (75% of the data) and test data (25% of the data). The training data were used to generate the model and the test data were used to verify model accuracy (Fielding, 2002). I checked the sensitivity of SDMs using the area under the curve (AUC) of the receiver operating characteristic (ROC). I used the threshold of equal sensitivity (omission errors) and specificity

(commission errors) to classify the output map from Maxent (Pearson, 2008; Hijmans et al.,

2015), and also to transform them into binary raster layers with presence and absence. In order to generate a climatic stability surface, the binary layers derived from SDMs of each time period were overlaid and overlap areas among all different time periods projections were checked, as a result I created an output a map with values ranging from 0 (not predicted to be present during any time period) to 12 (predicted to be present in all time periods).

All SDMs were built with R (R Development Core Team, 2017) using the packages raster (Hijmans & Etten, 2012), rgdal (Bivand et al., 2015a), dismo (Hijmans et al., 2015), rJava

(Urbanek, 2015), maptools (Bivand et al., 2015b), phyloclim (Heibl & Calenge, 2018),

RColorBrewer (Neuwirth, 2014), rgeos (Bivand & Rundel, 2013), maps (Becker et al., 2016),

GISTools (Brunsdon & Chen, 2015).

Analysis Molecular of Variance (AMOVA)

I performed a hierarchical analysis molecular of variance (Excoffier et al., 1992;

Grünwald & Hoheisel, 2007), to assess the evidence for the presence of genetic variance in my

SNPs data between populations and climatic stability, and between populations and geomorphologic processes (plateau vs. valley). According to Ab’Sáber (1983), the Brazilian

68

central plateau ranges from 300m to 1100m, and some studies defined the plateaus are above

500m of altitude (Silva, 1997; Santos et al., 2014). Thus, I defined the populations that were in the plateaus by setting the altitude to a minimum threshold of 500 m, and under 500 m they were defined valley populations. In order to define the populations as occurring in an area defined as climatically stable, I used the climatic stability surface (generated in the species distribution modeling section of the methods), and only the populations that were in the areas with high climatic stability, i.e. with the presence of climatic projection in all time periods, were considered climatically stable. The significance of the AMOVA was evaluated by a Monte-Carlo test performing 999 permutations. I performed both analyses in R (R Development Core Team,

2017) using the package poppr (Kamvar et al., 2014) and ade4 (Dray & Dufour, 2007).

Maps

To understand the distribution of the three species, I created distribution maps using R (R

Development Core Team, 2017) with the packages raster (Hijmans & Etten, 2012), rgdal

(Bivand et al., 2015a), maptools (Bivand et al., 2015b), and RColorBrewer (Neuwirth, 2014), and QGIS software (QGIS Development Team, 2015).

Results

3RADseq processing

My genetic sampling intended to obtain 2 million read per sample, after demultiplexing and filtering low quality reads, averaged of ~1,6 million reads per sample for a total of 86,827of total prefiltered loci for Chiasmocleis centralis, ~2.1 million reads for a total of 147,252 of total

69

loci for Dendropsophus rubicundulus, and ~1 million reads for a total of 89,425 of total loci for

Physalaemus nattereri. These loci were filtered via the iPyRAD pipeline to exclude loci that had low sequence coverage within samples and low representation across samples. A total of 36,407 loci passed this filtration step for an average of 10,545 loci per sample for C. centralis, 68,251 total filtered loci and 11,374 loci on average per sample for D. rubicundulus, and 35,534 total filtered loci and 8,898 loci on average per sample for P. nattereri. After the alignment of the loci across samples, the final data matrices had between 10,171 and 3,895 loci depending on the amount missing data allowed and whether the outgroups were included or not for C. centralis,

6,868 and 1,519 loci for D. rubicundulus, and 7,841 and 4,455 loci for P. nattereri (50% and

20% respectively; Table 8). Two samples each of Hypsiboas albopunctatus, D. marmoratus, D. leucophyllatus, D. minutus, and D. jimi as outgroups of D. rubicundulus; Ctenophryne geayi,

Dermatonotus muelleri, and Elachistocleis ovalis as outgroups of C. centralis; Pleurodema diplolister, P. centralis, P. cuvieri, and Pseudopaludicula mystacalis as outgroups of P. nattereri.

Phylogenetics, Phylogeography and population structure

The phylogenetic relationships inferred for all three focal species with RAxML using varying amounts of missing data were very similar, differing primarily in levels of branch support. For subsequent comparative phylogeographic analyses I used the SNP matrices resulting from sequence data processing that allowed up to 50% missing data (for number of SNPs see

Table 6). The phylogenetic hypotheses recovered by RAxML analyses recovered clades that were geographically congruent with the genetic clusters identified by the STRUCTURE analyses

(results below) with high support >95% for Chiasmocleis centralis (Fig. 14) and Dendropsophus

70

rubicundulus (Fig. 15), and for Physalaemus nattereri the support was not so high (Fig. 16). The species tree generated in SVDquartets (unlinked loci, again allowing up to 50% missing data) recovered the same geographic clusters for with high support for C. centralis (Fig. 17), D. rubicundulus (Fig. 18), and P. nattereri (Fig. 19). The STRUCTURE results (unlinked loci with

20% missing data allowed) for each of the focal species recovered three population clusters (Fig.

20). For C. centralis and P. nattereri analysis of the mean log-likelihood revealed values that increased until reaching a peak for the number of distinct genetic populations (K) = 3, with the

DK statistic also peaking at 3 populations clusters (Evanno et al., 2005). However, for D. rubicundulus the mean log-likelihood reached its peak at K=5, though the DK statistic provided clear support for three, rather than five populations clusters (Evanno et al., 2005). This difference highlights the importance of employing the DK statistic, as it identifies values of K above which increases in marginal likelihood are insufficient to warrant acceptance of this increasingly fragmented picture of population structure. The geographic distribution of individuals and their assignment to recovered population clusters for each of the three species can be observed in

Figures 21, 22, and 23. Henceforth, population clusters will be discussed by referring to them with respect to their relative position among the other populations for a particular species: C. centralis – Central-East, Central-North and West (Fig. 21 and Table 5); D. rubicundulus –

Central, North and South (Fig. 22 and Table 6); P. nattereri – Central, North and West (Fig. 23 and Table 7).

There were few differences between the RAxML and the SVD quartets tree topology, for all species the populations clusters found in the STRUCTURE analysis were recovered in the

SVDquartets tree (Fig. 17, 18 and 19). However, some localities have a different placement in the species tree when compared to the RAxML, for Dendropsophus rubicundulus the population

71

from São Desidério, BA (sdesi), have a different placement in the SVDquartets tree but preserved the population clusters assigned and were considered irrelevant. However, the placement of the population clusters had some noteworthy differences for Physalaemus nattereri, as North clade was the oldest divergent clade in the maximum likelihood tree while in the coalescent species tree estimated the oldest divergent clade was the Central clade.

Divergence time and Demographic history

According to the results of momi2 for Chiasmocleis centralis, the time of divergence between West and Central-North was ~4.4 Mya, and the divergence between West and Central-

East was ~4.5 Mya. The migration model selected (from those illustrated in Fig. 13) supports bi- directional pulses of migration around ~434 Kya between population West and Central-East and

~350 Kya between population Central-East and West (Fig. 24 and Table 9). The model selected for Dendropsophus rubicundulus was different and the time of divergence of populations South and Central was ~1.2 Mya, and between populations Central and North was ~643 Kya. However, the time of divergence between populations South and Central seems to be around ~2.5 Mya, and between populations Central and North was ~500 Kya after performing the bootstrap analyses.

Bi-directional events of migration occurred around ~350 Kya and ~76 Kya between populations

Central and North (Fig. 24 and Table 9). Finally, for Physalaemus nattereri the divergence time between populations North and West was ~2.87 Mya and ~2.83 Mya for West and Central. The uni-directional migration pulses occurred between population West and Central around ~278

Kya and between West and North around ~20 Kya (Fig. 24 and Table 9). The test for introgression between populations showed the presence of admixture for all species because all values for the z-scores are different from zero (Table 10), the negative values indicate that

72

populations are more similar to the ancestor population and the positive values indicate that the populations are more different from the ancestor population, which indicates the level of differentiation between populations.

Table 9: Results of model selection using AIC for the six demographic models analyzed (Figure 6), the models in bold were the selected for the bootstrap analyses. Species Model AIC Delta AIC AIC weight No migration 113256.27 7619.46 0 Population size 105784.59 147.78 8.09e-033 Chiasmocleis Migration 1 105753.06 116.26 5.67e-026 centralis Migration 2 105636.80 0 1 Migration 3 106276.05 639.24 1.54e-139 Migration 4 105935.46 298.662 1.40e-065 No migration 238577.25 11275.12 0 Population size 228865.47 1563.34 0 Dendropsophus Migration 1 227302.13 0 1 rubicundulus Migration 2 228799.21 1497.08 0 Migration 3 228500.23 1198.10 6.84e-261 Migration 4 227915.65 613.51 5.96e-134 Physalaemus No migration 124190.96 6021.55 0 nattereri Population size 119230.72 1061.32 3.44e-231 Migration 1 118937.52 768.12 1.59 e-167 Migration 2 118649.10 479.70 6.81 e-105 Migration 3 118295.05 125.65 5.18 e-28 Migration 4 118169.40 0 1

Table 10: Results of the f3-statistic for all species. Species Expected Observed SD Z C. centralis 0.091 0.114 0.004 4.98 D. rubicundulus 0.0099 0.0038 0.002 -2.95 P. nattereri 0.013 0.018 0.002 2.70

Population genetic diversity and isolation by distance (IBD)

The populations distributed in the central region of the Cerrado savannas have lower values of genetic diversity, except for the population North of Dendropsophus rubicundulus

(Table 11). In addition, the values observed for inbreeding (Fis) were similar for all population and for all species, except for the population South of D. rubicundulus for which an Fis estimate of 0.6 suggests inbreeding, and thus, a small population size (Table 11). The observed fixation

73

index values (Fst) suggest that there is a moderate population structuring for all species (Table

12). A strong effect of isolation by distance was observed among sample localities in all species

(C. centralis: r=0.67, p-value <0.001; D. rubicundulus: r=0.49, p-value <0.001; P. nattereri: r=0.49, p-value <0.001).

Table 11: Population genetic diversity of each species, sample size (n), inbreeding coefficient (Fis) and average expected heterozygosity (Hs). Species Population n Fis Hs Chiasmocleis West 9 0.34 0.0921 centralis Central-East 34 0.30 0.0626 Central-North 48 0.30 0.0768 Dendropsophus South 26 0.60 0.0845 rubicundulus Central 68 0.32 0.0736 North 21 0.25 0.0529 Physalaemus West 58 0.30 0.0668 nattereri Central 64 0.32 0.0434 North 59 0.35 0.0850

Table 12: Pair-wise Fst values between populations for each species. Species Par-wise Fst Chiasmocleis Populations West Central-East centralis Central-East 0.3851 Central-North 0.2969 0.4302 Dendropsophus Populations South Central rubicundulus Central 0.3772 North 0.4732 0.2024 Physalemus Populations West North nattereri North 0.4024 Central 0.3448 0.3692

Species distribution modeling

The predicted distribution of Chiasmocleis centralis, Dendropsophus rubicundulus and

Physalaemus nattereri for the current climate was very closely matched the Cerrado current extent, also including some of the Atlantic Forest in the prediction. Such overlap in the prediction with the adjacent biome might be explained by the inclusion of data points that occur in the transitional zone between Cerrado and Atlantic Forest (Fig. 25, 26, and 27). The species

74

distribution model (SDM) generated using current climate data had a good performance predicting the distribution of all species with an AUC of 0.88 for C. centralis, 0.80 for D. rubicundulus, and 0.84 for P. nattereri using cross-validation of random blocks of the occurrence data. The projection of the model on past climatic conditions showed that during the Late

Pliocene (3.3 Mya), when the climate was drier and colder (Dolan et al., 2015), the species distributions models were likely similar to, but slightly broader than the predicted distribution for the current climate for Chiasmocleis centralis and Physalaemus nattereri. During the mid-

Pliocene (3.2-3 Mya), the climatic shifts are predicted to have promoted an expansion of the species distribution area for Chiasmocleis centralis and Physalaemus nattereri, probably due warmer climatic conditions paired with high levels of CO2, similar to present-day (Hill, 2015).

This Pliocene SDM predicts the greatest distributional extent of all models/time periods for

Chiasmocleis centralis and Physalaemus nattereri. During the Pleistocene (785 Kya), the dry and cold climate promoted the contraction of the distribution of Chiasmocleis centralis and

Physalaemus nattereri, apparently lasting until after the Last Glacial Maximum, followed by a recent expansion of the predicted species distributions during the Holocene (Fig. 25, 26, and 27).

Dendropsophus rubicundulus distribution had different responses to the climate shifts, a smaller distribution range was observed during the Pliocene when compared to the current climate prediction (Fig. 26). The largest distribution range occurred during the Last Glacial Maximum when compared to the current climate species distribution (Fig. 26). The climate stability surface showed a large stable area in the central area of the Cerrado, predicting this to be an area of refugial isolation for each species during the climatic shifts (Fig. 28, 29, and 30).

75

Analysis Molecular of Variance (AMOVA)

Analysis of molecular variance (AMOVA) did not identify patterns of genetic differentiation related to geomorphology (plateau vs. valley populations) nor climatic stability in any of the species, also the percentage of explained variation was not associated with it as well

(Table 13). However, there was genetic differentiation within genetic populations between geomorphological formations (plateau/valley), and extremes of climatic stability (P ³ 0.001)

(Table 13). For Chiasmocleis centralis, a total of 54.02% of the variance was associated with differences within populations, whereas only 50.83% of the variance was attributable to variation between geomorphologic formations within populations, and a total of -4.86% of the variance was associated with differences among geomorphologic formations showing absence of genetic structure between plateaus and valleys (Table 13). Also, a total of 57.64% of the variance was associated with differences within populations, whereas only 56.60% of the variance was attributable to variation in climatic stability within populations, and a total of -14.24% of the variance was associated with differences among climatic stability showing absence of genetic structure between stable and unstable (Table 13). For Dendropsophus rubicundulus, a total of

59.25% of the variance was associated with differences within populations, whereas 57.72% of the variance was attributable to variation between geomorphologic formations within populations, and a total of -16.98% of the variance was associated with differences among geomorphologic formations showing absence of genetic structure between plateaus and valleys

(Table 13). Also, a total of 71.02% of the variance was associated with differences within populations, whereas 57.14% of the variance was attributable to variation in climatic stability within populations, and a total of -22.37% of the variance was associated with differences among climatic stability showing absence of genetic structure between stable and unstable (Table 13).

76

For Physalemus nattereri, a total of 59.27% of the variance was associated with differences within populations, whereas only 55.34% of the variance was attributable to variation between geomorphologic processes within populations, and a total of 14.64% of the variance was associated with differences among geomorphologic formations showing absence of genetic structure between plateaus and valleys (Table 13). Also, a total of 59.08% of the variance was associated with differences within populations, whereas only 52.94% of the variance was attributable to variation between climatic stability within populations, and a total of -12.02% of the variance was associated with differences among climatic stability showing absence of genetic structure between stable and unstable (Table 13). Fixation indices (F) were moderate but were lowest between geomorphologic processes and between climatic stability for all species, the negative values of F-statistic indicate no degree of differentiation between them (Table 13).

77

Table 13: Analysis of molecular variance for each species, testing for differences in geomorphologic processes (geo) and climatic stability (climatic). The values in bold represent significant variation (P ³ 0.001). Source of variation df Sum of Mean Variation Phi-statistic (F) squares squares (%) Chiasmocleis centralis Between geo 1 3201.41 3201.41 -4.86 geo-total -0.04 Between populations within geo 2 4396.15 2198.07 50.83 populations-geo 0.48 Within populations 87 11530.45 132.53 54.02 populations-total 0.45 Total 90 19128.02 212.53 100 Between climatic 1 982.05 982.05 -14.24 climatic-total -0.14 Between populations within populations- 3 6761.63 2253.87 56.60 0.49 climatic climatic Within populations 86 11384.32 132.37 57.64 populations-total 0.42 Total 90 19128.02 212.53 100 Dendropsophus rubicundulus Between geo 1 555.01 555.01 -16.98 geo-total -0.16 Between populations within geo 4 2157.66 539.60 57.72 populations-geo 0.49 Within populations 109 4303.85 39.48 59.25 populations-total 0.40 Total 114 7016.53 61.54 100 Between climatic 1 185.40 185.40 -22.37 climatic-total -0.22 Between populations within populations- 4 2001.25 500.31 51.34 0.41 climatic climatic Within populations 109 4829.87 44.31 71.02 populations-total 0.28 Total 114 7016.53 61.54 100 Physalaemus nattereri Between geo 1 1084.17 1084.17 -14.64 geo-total -0.14 Between populations within geo 4 13308.84 3327.21 55.37 populations-geo 0.48 Within populations 175 20875.17 119.28 59.27 populations-total 0.40 Total 180 35268.20 195.93 100 Between climatic 1 1511.08 1511.08 -12.02 climatic-total -0.12 Between populations within populations- 3 12378.03 4126.01 52.94 0.47 climatic climatic Within populations 176 21379.09 121.47 59.08 populations-total 0.40 Total 180 35268.20 195.93 100

Discussion

I recovered a strong spatial genetic structure for all species, all of which have three well- supported genetic clusters and geographically structured populations that are widespread in the

Neotropical savannas. These genetic clusters are roughly distributed in the west, central and north of the Cerrado savannas for all species, which is in agreement with recently published phylogeographic studies of amphibian species that also found three (Hypsiboas albopunctata) or

78

four (H. lundii) population clusters distributed in the west, central west, central and southeast of the Cerrado (Prado et al., 2012; Vasconcellos et al., 2019). A similar pattern was also observed for the lizard Micrablepharus atticulus, with three groups distributed roughly in the west, central and south of the Cerrado (Santos et al., 2014). The similarity observed for these different species of amphibians and one lizard could be an outcome of the landscape evolution during the Pliocene and Pleistocene, and the climatic fluctuations of the Quaternary. In addition, phylogenetic relationships (maximum likelihood and coalescent species tree) of population clusters were very similar across species, recovering the same phylogeographic clades with minor differences in the placement of localities. However, the placement of the population clusters had some noteworthy differences for Physalaemus nattereri, the North clade was the oldest divergent clade in the maximum likelihood tree while in the coalescent estimated species tree the oldest divergent clade was the Central clade, showing the variation between the methods used can impact the interpretation of the results and the demography analysis.

In recent studies, the population clusters identified that are distributed in the West/South- western region are usually the oldest divergent clade in the phylogenetic tree, sister to the populations in the Central/Central-east, and the populations in the South-east are recent lineages

(Prado et al., 2012; Santos et al., 2014; Guarnizo et al., 2016; Miranda et al., 2019; Vasconcellos et al., 2019). However, the evolutionary pattern found in my study is mostly southwestern to north of the Cerrado, Central or South-western clades are the oldest divergent clade in the tree, followed by West clades, and North clades are more recent lineages. This pattern corroborates the prediction that the initial divergence of ancestral populations of these species occurred in central of Cerrado. There is an evidence of connections between rain forests with dry open vegetations for Neotropical savannas vertebrates species (Costa, 2003; Lynch Alfaro et al., 2012;

79

Fouquet et al., 2014; Gehara et al., 2014). The center of origin of theses vertebrates was the

Amazon Forest, with dispersion and speciation at Atlantic Forest, followed by dispersal and speciation to Chaco and Neotropical savannas (Bolivian and Brazilian) (Costa, 2003; Lynch

Alfaro et al., 2012; Fouquet et al., 2014; Gehara et al., 2014). Thus, this diversification pattern observed with species dispersing from Chaco and Bolivian savannas to the south-western part of

Cerrado, could explain the distribution pattern observed for the populations.

The niche modeling confirmed the cooling tendency of the earth’s surface during

Pliocene with contraction of the predicted distribution of Chiasmocleis centralis and

Physalaemus nattereri. However, Dendropsophus rubicundulus had an expansion of the distribution with a cooler environment. The climatic shifts promoted lowering of the sea level with the colder climate, and dryer and cooler conditions promoted contraction and expansion of the lowland forests and open vegetations, respectively (Haffer, 1979). I could observe the distribution shifts for Chiasmocleis centralis and Physalaemus nattereri, with reduction of the predicted distribution around 3.3 Mya when the climate when was drier and colder, and somewhat similar to the current climate (Dolan et al., 2015). Followed by expansion around 3.2 -

3 Mya when the climatic conditions became warmer than the current climatic conditions (Hill,

2015). After that, a dramatic decrease in the distribution when the climate became colder and dryer until reaching the Last Glacial Maximum (21 Kya), and finally with the increase of the

Earth’s temperature reaching the current predict distribution. The distribution shifts observed for

C. centralis, and P. nattereri, is in agreement with what is expected for the Cerrado vegetations, i.e. contraction when it is cold and dry vs. expansion when is warm and wet, reaching a narrow distribution in the LGM (21 Kya) (Werneck et al., 2012b). Likewise, it is in agreement with the predicted distribution shifts of other Cerrado species (Santos et al., 2014; Miranda et al., 2019;

80

Vasconcellos et al., 2019). Dendropsophus rubicundulus had different distributional responses that is in disagreement with the Cerrado vegetations and Cerrado species distributions. However, these studies only used fewer climate data from the Upper Pleistocene to current climate (four time periods total), while my predictions included data from Late Pliocene with twelve different time periods. My species distribution modeling has more climate periods, which increased the complexity of the species distribution modeling, making difficult to compare with what happened for the Cerrado species during the Pliocene and middle Pleistocene. Furthermore, the time of diversification between populations of Chiasmocleis centralis, Dendropsophus rubicundulus and Physalaemus nattereri occurred approximately 4.5, 2.5, and 2.8 Mya, respectively, and unfortunately there is no climatic data available in the same time frames for

SDM. Thus, I cannot make precise inferences about the impacts of the climatic shifts on the population’s divergence. However, the time of divergence between populations occurred during important geomorphological events in South America, and were also important for the diversification of other amphibian species (Maciel et al., 2010; Prado et al., 2012).

In combination with the climatic shifts, the geomorphological events during the late

Miocene and early Pliocene (~ 5 Mya) promoted dispersal routes from forested biomes to the dry open vegetation of the Neotropical savannas. The estimated times of the divergence of populations within Chiasmocleis centralis, Dendropsophus rubicundulus and Physalaemus nattereri suggest the effect of the geomorphological events may have contributed to the isolation of populations during the Pliocene and Pleistocene. The divergence of the populations for all species occurred during the Pliocene (~4.5 Mya) and Pleistocene (~2.8 and ~2.5 Mya), with migration events during the middle Pleistocene and Holocene, with one congruent migration event in the middle Pleistocene (~350 Kya). These results suggest that the further upwarping of

81

the Brazilian shield that resulted in the subsidence of the peripheral depression (Chaco and

Cerrado) during the Pliocene and early Pleistocene (5 - 2 Mya) (Mendes & Petri, 1971; Haffer,

1974; Colli, 2005; Morrone, 2014) were important geological events promoting diversification of

C. centralis, D. rubicundulus and P. nattereri. The impacts of the same geomorphological changes during the Pliocene e Pleistocene in the diversification were observed for amphibians and lizard species in the Cerrado savannas (Maciel et al., 2010; Prado et al., 2012; Santos et al.,

2014; Miranda et al., 2019). Another geomorphological event that could have impacted the diversification of amphibians in the Cerrado is the formation of the Paraná river basin during the

Miocene and Pliocene (Brea & Zucol, 2011). This was suggested to be important for the

Chacoan species by Gallardo (1979) and my study included samples of P. nattereri from the

Chaco biome. It is known that the inland seas in the south of South America (Paranaense Sea) disappeared during the late Miocene (~11 - 5 Mya) and there is no evidence of marine transgressions after that (Cione et al., 2011). However, the landscape could remain unforested and uninhabitable for thousands or millions of years, having some small salty inland lakes or salty wetlands, that did allow the dispersal because of harsh environmental conditions during the

Pliocene. Thus, knowing the connections between Amazon and Atlantic Forest with Cerrado and

Chaco were also observed for other vertebrate species (Costa, 2003; Lynch Alfaro et al., 2012;

Fouquet et al., 2014; Gehara et al., 2014), and it is possible that the formation of the Paraná basin had a great impact in the diversification of Cerrado’s herpetofauna.

The demographic models selected included migration events between populations, and only moderate genetic structuring was observed, confirming that there are no major geographic barriers to gene flow. My demographic models did not test for population expansion or bottlenecks, so these results cannot evaluate precisely the effects of climatic fluctuations on the

82

size of populations. However, these results do show a strong effect of isolation by distance between populations, and the climatic fluctuations and the geomorphological formations alone were not significant predictors of genetic variation. Instead, observed population structure explains the majority of genetic variation observed, and my predictions of higher levels of genetic variation in areas of climatic stability and on the plateaus (vs. valleys) were not supported. The observed absence of a relationship between geological formation and genetic diversity is in agreement with at least one other published study of a Cerrado species (Santos et al., 2014), as is the absence of any difference in genetic diversity between populations distributed in the climatically stable vs. unstable areas (Miranda et al., 2019). According to my species distribution modeling, there is a large climatically stable area that covers most of the central

Cerrado, and the species modeling for past periods showed shifts in the climatic suitability towards west to southeast during the Quaternary, which is in agreement with a recent study on

Hypsiboas lundii in the Cerrado (Vasconcellos et al., 2019). According Vasconcellos et al.

(2019), the climatic suitability might have forced distributional shifts from west to east direction, and the climatic instability promoted range shifts that could have impacted negatively the gene flow promoting differentiation between populations (Vasconcellos et al., 2019). Thus, the diversification pattern observed for Hypsiboans lundii could also explain the diversification of C. centralis, D. rubicundulus, and P. nattereri.

Conclusion

The results are aligned with diversification patterns described by recent research, which reported three to four population clusters and diversification directionality from west to east

(Prado et al., 2012; Santos et al., 2014; Miranda et al., 2019; Vasconcellos et al., 2019). For

83

Chiasmocleis centralis, Dendropsophus rubicundulus and Physalaemus nattereri, there are three populations widely distributed in the Cerrado. However, the directionality of the population’s diversification found in my study is mostly southwestern to north of the Neotropical savannas, conflicting with the recent research (Prado et al., 2012; Santos et al., 2014; Miranda et al., 2019;

Vasconcellos et al., 2019). The time of divergence of the populations indicated that geological events during the Neogene (~ 5 – 2 Mya) were important for the divergence of the populations.

The species distribution modeling for Chiasmocleis centralis, Dendropsophus rubicundulus and

Physalaemus nattereri showed the effect of climatic changes on the distributional range shifts, with a retraction of species range from west to east after the Last Glacial Maximum (21 Kya), followed by expansion from east to north in Holocene (~ 11 Kya). The demographic models indicate the presence migration between populations, indicating that there are no important geographic barriers for gene flow. However, the demographic models did not include population expansion or bottlenecks according the predicted distribution shifts found in the niche modeling, but these events should be incorporated. Thus, it is possible that isolation by distance also promoted the differentiation among the populations, in combination with the climatic fluctuations during the Quaternary. For future directions, the implementation of an integrative comparative phylogeography analysis with C. centralis, D. rubicundulus and P. nattereri is necessary to evaluate the synchronicity of the demographic models and propose a general pattern of diversification for the amphibians of Cerrado.

84

FIGURES

B A C B A C B A C

No migration Population size Migration 1

B A C B A C B A C

Migration 2 Migration 3 Migration 4

Figure 13. Demography models tested using momi2. (1) divergence with no migration not considering the population size; (2) divergence with no migration considering the estimated population size; (3) divergence with bidirectional migration between populations A and B, considering the estimated population size; (4) divergence with bidirectional migration between populations B and C, considering the estimated population size; (5) divergence with unidirectional migration between populations C and B, and B and A, considering the estimated population size; (6) divergence with unidirectional migration between populations B and A, B and C, considering the estimated population size.

85

Figure 14. Maximum likelihood tree for Chiasmocleis centralis obtained from RAxML. Populations are represented by the colors: purple = West; orange = Central-East; and blue = Central-North.

86

Figure 15. Maximum likelihood tree for Dendropsophus minutus obtained from RAxML. Populations are represented by the colors: purple = South; orange = Central; and blue = North.

87

Figure 16. Maximum likelihood tree for Physalaemus nattereri obtained from RAxML. Populations are represented by the colors: purple = West; orange = Central; and blue = North.

88

Figure 17. Coalescent model species tree for Chiasmocleis centralis obtained from SVDquartets. Populations are represented by the colors: purple = West; orange = Central-East; and blue = Central-North.

89

Figure 18. Coalescent model species tree for Dendropsophus minutus obtained from SVDquartets. Populations are represented by the colors: purple = South; orange = Central; and blue = North.

90

Figure 19. Coalescent model species tree for Physalaemus nattereri obtained from SVDquartets. Populations are represented by the colors: purple = West; orange = Central; and blue = North.

91

Figure 20. The three genetic clusters found, mean likelihood and DK for (A, b and c) Dendropsophus rubicundulus, (D, c and e) Chiasmocleis centralis, and (G, h and i) Physalaemus nattereri.

92

Figure 21. Distribution map of Chiasmocleis centralis, in red the Cerrado range and number are the localities (Table 3), the gray scale is elevation (m).

93

Figure 22. Distribution map of Dendropsophus rubicundulus, in red the Cerrado range and number are the localities (Table 4), the gray scale is elevation (m).

94

Figure 23. Distribution map of Physalaemus nattereri, in red the Cerrado range and number are the localities (Table 5), the gray scale is elevation (m).

95

Figure 24. The demographic models selected for (A) Chiasmocleis centralis, (B) Dendropsophus rubicundulus, and (C) Physalaemus nattereri.

96

Figure 25. Species distribution modeling for Chiasmocleis centralis under current climatic condition and projected for 12 different past climate data using Maxent.

97

Figure 26. Climatic stability surface using all time periods for Chiasmocleis centralis, values ranging from 0 (no present) to 12 (present in all time periods), areas with high climatic stability have high values.

98

Figure 27. Species distribution modeling for Dendropsophus rubicundulus under current climatic condition and projected for 12 different past climate data using Maxent.

99

Figure 28. Climatic stability surface using all time periods for Dendropsophus rubicundulus, values ranging from 0 (no present) to 12 (present in all time periods), areas with high climatic stability have high values.

100

Figure 29. Species distribution modeling for Physalaemus nattereri under current climatic condition and projected for 12 different past climate data using Maxent.

101

Figure 30. Climatic stability surface using all time periods for Physalaemus nattereri, values ranging from 0 (no present) to 12 (present in all time periods), areas with high climatic stability have high values.

102

LIST OF REFERENCES

103

Ab’Saber, A. N. 1977. Os domínios morfoclimáticos da América do Sul. Primeira Aproximação. Geomorfologia 52:1–21. Ab’Sáber, A. N. 1983. O domínio dos Cerrados. Revista Servidor Público, Brasília-DF 3:41–55. Affonso, I. de P., E. G. Cafofo, R. L. Delariva, F. H. Oda, L. C. Karling, & R. Lourenço-de- Moraes. 2014. List of anurans (Amphibia: Anura) from the rural zone of the municipality of Maringá, Paraná state, southern Brazil. Check List 10:878–882. Aljanabi, S. M., & I. Martinez. 1997. Universal and rapid salt-extraction of high quality genomic DNA for PCR-based techniques. Nucleic Acids Research 25:4692–4693. Amirisetty, S., G. K. K. Hershey, & T. M. Baye. 2012. AncestrySNPminer: A bioinformatics tool to retrieve and develop ancestry informative SNP panels. Genomics 100:57–63. Annunziata, B. B., W. M. Fontenele, I. S. Castro, R. F. Gurgel, & K. A. do Nascimento. 2007. Amphibia, Anura, Hylidae, Dendropsophus rubicundulus: distribution extension. Check List 3:242–243. Aquino, L., G. Colli, S. Reichle, D. Silvano, & N. Scott. 2004a. Chiasmocleis albopunctata. The IUCN Red List of Threatened Species. [Online.] Available at www.iucnredlist.org. Aquino, L., S. Reichle, D. Silvano, & N. Scott. 2004b. Eupemphix nattereri. The IUCN Red List of Threatened Species. [Online.] Available at www.iucnredlist.org. Avise, J. C., J. Arnold, R. M. Ball, E. Bermingham, T. Lamb, J. E. Neigel, C. A. Reeb, & N. C. Saunders. 1987. INTRASPECIFIC PHYLOGEOGRAPHY: The Mitochondrial DNA Bridge Between Population Genetics and Systematics. Annual Review of Ecology and Systematics 18:489–522. Avise, J. C., B. W. Bowen, & F. J. Ayala. 2016. In the light of evolution X: Comparative phylogeography. Proceedings of the National Academy of Sciences USA 113:7957–7961. Baele, G., P. Lemey, T. Bedford, A. Rambaut, M. A. Suchard, & A. V Alekseyenko. 2012. Improving the accuracy of demographic and molecular clock model comparison while accommodating phylogenetic uncertainty. Molecular Biology and Evolution 29:2157–2167. Baird, N. A., P. D. Etter, T. S. Atwood, M. C. Currey, A. L. Shiver, Z. A. Lewis, E. U. Selker, W. A. Cresko, & E. A. Johnson. 2008. Rapid SNP discovery and genetic mapping using sequenced RAD markers. PLoS ONE 3:e3376. Barley, A. J., A. N.-M. de Oca, T. W. Reeder, N. L. Manríquez-Morán, J. C. A. Monroy, O. Hernández-Gallegos, & R. C. Thomson. 2018. Complex patterns of hybridization and introgression across evolutionary timescales in Mexican whiptail lizards (). Molecular Phylogenetics and Evolution 132:284–295. Becker, R. A., A. R. Wilks, R. Brownrigg, T. P. Minka, & A. Deckmyn. 2016. maps: Draw Geographical Maps. R package version 3.1.0:http://cran.r-project.org/package=maps. Behling, H. 1995. A high resolution Holocene pollen record from Lago do Pires, SE Brazil: vegetation, climate and fire history. Journal of Paleolimnology 14:253–268. Bermingham, E., & C. Moritz. 1998. Comparative phylogeography: concepts and applications. Molecular Ecology:367–369. Besnier, F., & K. A. Glover. 2013. ParallelStructure: A R Package to Distribute Parallel Runs of the Population Genetics Program STRUCTURE on Multi-Core Computers. PLoS ONE

104

8:1–5. Bigarella, J. J., & D. de Andrade-Lima. 1982. Paleoenvironmental changes in Brazil. In: Biological diversification in the Tropics. G. T. Prance (ed.). Columbia University Press, New York, pp. 27–40. Bivand, R., T. Keitt, B. Rowlingson, E. Pebesma, M. Sumner, R. Hijmans, & E. Rouault. 2015a. rgdal: Bindings for the geospatial data abstraction library. R package version 0.9-2. Bivand, R., N. Lewin-Koh, E. Pebesma, E. Archer, A. Baddeley, H.-J. Bibiko, S. Brey, J. Callahan, G. Carrillo, S. Dray, D. Forrest, et al. 2015b. maptools: Tools for reading and handling spatial objects. R package version 0.8-34. Bivand, R., & C. Rundel. 2013. rgeos: interface to geometry engine-open source (GEOS). R package version 0.3-2. Bokermann, W. C. A. 1952. Microhylidae de coleção do Departamento de Zoologia (Amphibia-Anura). Papéis Avulsos de Zoologia, São Paulo 10:271–292. Bouckaert, R., J. Heled, D. Kühnert, T. Vaughan, C.-H. Wu, D. Xie, M. A. Suchard, A. Rambaut, & A. J. Drummond. 2014. BEAST 2: A Software Platform for Bayesian Evolutionary Analysis. PLoS Computational Biology 10:e1003537. Brasileiro, C. A., E. M. Lucas, H. M. Oyamguchi, M. T. C. Thomé, & M. Dixo. 2008. Anurans, Northern Tocantins River Basin, states of Tocantins and Maranhão, Brazil. Check List 4:185–197. Brasileiro, C. A., R. J. Sawaya, M. C. Kiefer, & M. Martins. 2005. Amphibians of an open Cerrado fragment in southeastern Brasil. Biota Neotropica 5:1–17. Brea, M., & A. F. Zucol. 2011. The Paraná-Paraguay Basin: Geology and Paleoenvironments. In: Historical Biogeography of Neotropical Freshwater Fishes. J. Albert (ed.). University of California Press, pp. 69–87. Brown, J. H., & M. V Lomolino. 2006. Biogeografia. In.: 2nd edition. FUNPEC Ed., Ribeirão Preto. Brown, J. L., D. J. Hill, A. M. Dolan, A. C. Carnaval, & A. M. Haywood. 2018. Paleoclim, high spatial resolution paleoclimate surfaces for global land areas. Scientific Data 5:1–9. Brown, K. S., & D. R. Gifford. 2002. Lepidoptera in the Cerrado landscape and the conservation of vegetation, soil, and topographical mosaics. In: The Cerrados of Brazil: Ecology and natural history of a Neotropical savanna. P. S. Oliveira and R. J. Marquis (eds.). Columbia University Press, New York, NY, pp. 201–222. Brumfield, R. T., & S. V Edwards. 2007. Evolution into and out of the Andes: a Bayesian analysis of historical diversification in Thamnophilus antshrikes. Evolution 61:346–367. Brunsdon, C., & H. Chen. 2015. GISTools: Some further GIS capabilities for R. R package version 0.7-4. Brusquetti, F., & E. O. Lavilla. 2006. Lista comentada de los anfibios de Paraguay . Cuadernos de Herpetología 20:33–79. Bryant, D., R. Bouckaert, J. Felsenstein, N. A. Rosenberg, & A. Roychoudhury. 2012. Inferring species trees directly from biallelic genetic markers: Bypassing gene trees in a full coalescent analysis. Molecular Biology and Evolution 29:1917–1932.

105

Caramaschi, U., & C. A. G. da Cruz. 1997. Redescription of Chiasmocleis albopunctata (Boettger) and description of a new species of Chiasmocleis (Anura: Microhylidae). Herpetologica 53:259–268. Cardoso, A. J., & J. M. E. Vielliard. 1984. Caracterização bio-acústica da população topotipica de rubicundula (Amphibia, Anura). Revista Brasileira de Zoologia 2:423–426. Carnaval, A. C., M. J. Hickerson, C. F. B. Haddad, M. T. Rodrigues, & C. Moritz. 2009. Stability predicts genetic diversity in the Brazilian Atlantic Forest hotspot. Science 323:785–789. Cei, J. M. 1980. Amphibians of Argentina. Monitore Zoologico Italiano (N.S.) Monografia:1– 609. Chifman, J., & L. Kubatko. 2014. Quartet inference from SNP data under the coalescent model. Bioinformatics 30:3317–3324. Cione, A. L., M. A. Cozzuol, M. T. Dozo, & C. A. Hospitaleche. 2011. Marine vertebrate assemblages in the southwest Atlantic during the Miocene. Biological Journal of the Linnean Society 103:423–440. Collevatti, R., S. Rabelo, J. Nabout, & J. Diniz-Filho. 2011. Tracing the history of South- American neotropical savannas and seasonally dry forests evidences from comparative phylogeography. BMC Proceedings 5:I3. Colli, G. R. 2005. As origens e a diversificação da herpetofauna do Cerrado. In: Cerrado: Ecologia, Biodiversidade e Conservação. A. Scariot, J. C. Sousa-Silva and J. M. Felfili (eds.). Ministério do Meio Ambiente, Brasília, DF, pp. 247–264. Colli, G. R., R. P. Bastos, & A. F. B. Araujo. 2002. The Character and Dynamics of the Cerrado Herpetofauna. In: The Cerrados of Brazil: Ecology and Natural History of a Neotropical Savanna. P. S. Oliveira and R. J. Marquis (eds.). Columbia University Press, New York, pp. 223–241. Colli, G., S. Reichle, & D. Silvano. 2004. Dendropsophus rubicundulus. The IUCN Red List of Threatened Species. [Online.] Available at www.iucnredlist.org. Costa, L. P. 2003. The historical bridge between the Amazon and the Atlantic Forest of Brazil: a study of molecular phylogeography with small mammals. Journal of Biogeography 30:71– 86. Cruz, C. A. G., U. Caramaschi, & E. Izecksohn. 1997. The genus Chiasmocleis Méhelÿ, 1904 (Anura, Microhylidae) in the Atlantic Rain Forest of Brazil, with description of three new species. Alytes 15:49–71. Damasceno, R., M. L. Strangas, A. C. Carnaval, M. T. Rodrigues, & C. Moritz. 2014. Revisiting the vanishing refuge model of diversification. Frontiers in Genetics 5:1–12. Davey, J. W., J. L. Davey, M. L. Blaxter, & M. W. Blaxter. 2010. RADSeq: next-generation population genetics. Briefings in Functional Genomics 9:416–423. Dolan, A. M., A. M. Haywood, S. J. Hunter, J. C. Tindall, H. J. Dowsett, D. J. Hill, & S. J. Pickering. 2015. Modelling the enigmatic Late Pliocene Glacial Event - Marine Isotope Stage M2. Global and Planetary Change 128:47–60. Domingos, F. M. C. B., R. J. Bosque, J. Cassimiro, G. R. Colli, M. T. Rodrigues, M. G.

106

Santos, & L. B. Beheregaray. 2014. Out of the deep: Cryptic speciation in a Neotropical (, ) revealed by species delimitation methods. Molecular Phylogenetics and Evolution 80:113–124. Domingos, F. M. C. B., G. R. Colli, A. Lemmon, E. M. Lemmon, & L. B. Beheregaray. 2017. In the shadows: Phylogenomics and coalescent species delimitation unveil cryptic diversity in a Cerrado endemic lizard (Squamata: Tropidurus). Molecular Phylogenetics and Evolution 107:455–465. Dray, S., & A.-B. Dufour. 2007. The ade4 Package: Implementing the Duality Diagram for Ecologists. Journal of Statistical Software 1. Duellman, W. E. 1979. The South American herpetofauna: a panoramic view. In: The South American Herpetofauna: Its Origen, Evolution, and Dispersal. W. E. Duellman (ed.). University of Kansas Printing Service, Lawrence, pp. 1–28. Earl, D., & B. vonHoldt. 2012. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genetics Resources 4:359–361. Eaton, D. A. R. 2014. PyRAD: assembly of de novo RADseq loci for phylogenetic analyses. Bioinformatics 30:1844–1849. Eiten, G. 1972. The Cerrado vegetation of Brazil. Botanical Review 38:201–341. Emerson, K. J., C. R. Merz, J. M. Catchen, P. A. Hohenlohe, W. A. Cresko, W. E. Bradshaw, & C. M. Holzapfel. 2010. Resolving postglacial phylogeography using high- throughput sequencing. Proceedings of the National Academy of Sciences USA 107:16196– 16200. Evanno, G., S. Regnaut, & J. Goudet. 2005. Detecting the number of clusters of individuals using the software STRUCTURE: A simulation study. Molecular Ecology 14:2611–2620. Excoffier, L., P. E. Smouse, & J. M. Quattro. 1992. Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics 131:479–491. Faivovich, J., C. F. B. Haddad, P. C. A. Garcia, D. R. Frost, J. A. Campbell, & W. C. Wheeler. 2005. Systematic review of the frog family hylidae, with special reference to : phylogenetic analysis and taxonomic revision. Bulletin of the American Museum of Natural History:1–240. Fielding, A. H. 2002. What are the appropriate characteristics of an accuracy measure?. In: Predicting Species Occurrences: Issues of Accuracy and Scale. J. M. Scott, P. J. Heglunf, M. M. L., J. B. Haufler, M. G. Raphael, W. A. Wall and F. B. Samson (eds.). Island Press, Washington, DC, USA, pp. 271–280. Fordham, D. A., F. Saltré, S. Haythorne, T. M. L. Wigley, B. L. Otto-Bliesner, K. C. Chan, & B. W. Brook. 2017. PaleoView: a tool for generating continuous climate projections spanning the last 21 000 years at regional and global scales. Ecography 40:1348–1358. Forlani, M. C., P. H. Valdujo, D. Pavan, A. O. Maciel, & P. l. V Peloso. 2011. Review of the geographical distribution of the white-spotted frog Chiasmocleis albopunctata (Boettger, 1885) (Anura, Microhylidae). Biota Neotropica 11.

107

Fouquet, A., C. Santana Cassini, C. Fernando Baptista Haddad, N. Pech, & M. Trefaut Rodrigues. 2014. Species delimitation, patterns of diversification and historical biogeography of the Neotropical frog genus Adenomera (Anura, Leptodactylidae). Journal of Biogeography 41:855–870. Françoso, E., A. R. Zuntini, A. C. Carnaval, & M. C. Arias. 2016. Comparative phylogeography in the Atlantic forest and Brazilian savannas: Pleistocene fluctuations and dispersal shape spatial patterns in two bumblebees. BMC Evolutionary Biology 16:1–16. Frost, D. R. 2017. Amphibian Species of the World: an Online Reference. Version 6.0 (11/22/2017). American Museum of Natural History, New York, USA. Fujita, M. K., A. D. Leaché, F. T. Burbrink, J. A. McGuire, & C. Moritz. 2012. Coalescent- based species delimitation in an integrative taxonomy. Trends in Ecology and Evolution 27:480–488. Gallardo, J. M. 1979. Composición, distribución y origen de la herpetofauna chaqueña. In: The South American Herpetofauna - Its origin, Evolution and Dispersal. W. E. Duellman (ed.). Monographs. The University of Kansas, Lawrence, Kansas, pp. 299–307. Gamble, T., G. R. Colli, M. T. Rodrigues, F. P. Werneck, & A. M. Simons. 2012. Phylogeny and cryptic diversity in (Phyllopezus; Phyllodactylidae; Gekkota) from South America’s open biomes. Molecular Phylogenetics and Evolution 62:943–953. Gehara, M., A. J. Crawford, V. G. D. Orrico, A. Rodríguez, S. Lötters, A. Fouquet, L. S. Barrientos, F. Brusquetti, I. De La Riva, R. Ernst, G. G. Urrutia, et al. 2014. High levels of diversity uncovered in a widespread nominal taxon: Continental phylogeography of the neotropical tree frog dendropsophus minutus. PLoS ONE 9:e103958. Giaretta, A. A., & K. G. Facure. 2006. Terrestrial and communal nesting in Eupemphix nattereri (Anura, Leiuperidae): interactions with predators and pond structure. Journal of Natural History 40:2577–2587. Giselda, D. 2006. Observations on the Southern Cerrados and their relationship with the core area. In: Neotropical Savannas and Seasonally Dry Forests. CRC Press, pp. 67–77. Giugliano, L. G., R. G. Collevatti, & G. R. Colli. 2007. Molecular dating and phylogenetic relationships among (Squamata) inferred by molecular and morphological data. Molecular Phylogenetics and Evolution 45:168–179. Glenn, T., N. Bayona-Vásquez, T. Kieran, T. Pierson, S. Hoffberg, P. Scott, K. Bentley, J. J. Finger, P. Watson, S. Louha, N. Troendle, et al. 2017. Adapterama III: Quadruple- indexed, triple-enzyme RADseq libraries for about $1USD per Sample (3RAD). bioRxiv 205799:1–34. Goudet, J. 2005. hierfstat, a package for r to compute and test hierarchical F-statistics. Molecular Ecology Notes 5:184–186. Gross, M., M. I. F. Ramos, & W. E. Piller. 2016. A minute ostracod (Crustacea: Cytheromatidae) from the Miocene Solimões Formation (western Amazonia, Brazil): Evidence for marine incursions?. Journal of Systematic Palaeontology 14:581–602. Grünwald, N. J., & G.-A. Hoheisel. 2007. Hierarchical Analysis of Diversity, Selfing, and Genetic Differentiation in Populations of the Oomycete Aphanomyces euteiches . Phytopathology 96:1134–1141.

108

Guarnizo, C. E., F. P. Werneck, L. G. Giugliano, M. G. Santos, J. Fenker, L. Sousa, A. B. D’Angiolella, A. R. dos Santos, C. Strüssmann, M. T. Rodrigues, T. F. Dorado- Rodrigues, et al. 2016. Cryptic lineages and diversification of an endemic anole lizard (Squamata, Dactyloidae) of the Cerrado hotspot. Molecular Phylogenetics and Evolution 94:279–289. Haffer, J. 1974. Geologic and climatic history of tropical South America. In: Avian speciation in tropical South America. Raymond A. Paynter (ed.). Publication of the Nuttall Ornithological Club, NO. 14, Cambridge, Massachusetts, pp. 121–142. Haffer, J. 1979. Quaternary biogeography of tropical lowland South America. In: The South American Herpetofauna: Its Origen, Evolution, and Dispersal. W. E. Duellman (ed.). The University of Kansas, Lawrence, Kansas, pp. 107–140. Haga, I. A., F. S. de Andrade, D. P. Bruschi, S. M. Recco-Pimentel, & A. A. Giaretta. 2017. Unrevealing the leaf frogs Cerrado diversity: A new species of Pithecopus (Anura, Arboranae, Phyllomedusidae) from the Mato Grosso state, Brazil. PLoS ONE 12:e0184631. Heibl, C., & C. Calenge. 2018. phyloclim: Integrating Phylogenetics and Climatic Niche Modeling. R package version0.9.5. Henriques, D., M. Parejo, A. Vignal, D. Wragg, A. Wallberg, M. T. Webster, & M. A. Pinto. 2018. Developing reduced SNP assays from whole-genome sequence data to estimate introgression in an organism with complex genetic patterns, the Iberian honeybee ( Apis mellifera iberiensis ). Evolutionary Applications 11:1270–1282. Heyer, W. R. 1988. On frog distribution patterns east of the Andes. In: Proceedings of the workshop on neotropical distribution patterns. E. Vanzolini and W. R. Heyer (eds.). Academia Brasileira de Ciências, Rio de Janeiro. Hijmans, R. J. 2005. Very high resolution interpolated climate surfaces for global land areas. International Journal of Climatology 25:1965–1978. Hijmans, R. J., & J. van Etten. 2012. raster: Geographic analysis and modeling with raster data. R package version 2.0-12. Hijmans, R. J., S. Phillips, J. Leathwick, & J. Elith. 2015. dismo: Species distribution modeling. R package version 1.0-12. Hill, D. J. 2015. The non-analogue nature of Pliocene temperature gradients. Earth and Planetary Science Letters 425:232–241. Hoffmann, F. G., & R. J. Baker. 2003. Comparative phylogeography of short-tailed bats (Carollia: Phyllostomidae). Molecular Ecology 12:3403–3414. Humphries, C. J., & L. R. Parenti. 1999. Cladistic biogeography: interpreting patterns of plant and animal distributions. In.: 2nd edition. Oxford University Press. Jansen, M., R. Bloch, A. Schulze, & M. Pfenninger. 2011. Integrative inventory of Bolivia’s lowland anurans reveals hidden diversity. Zoologica Scripta 40:567–583. Jepson, W. 2005. A disappearing biome? Reconsidering land-cover change in the Brazilian savanna. Geographical Journal 171:99–111. Jombart, T. 2008. adegenet: a R package for the multivariate analysis of genetic markers. Bioinformatics 24:1403–1405.

109

Kamm, J. A., J. Terhorst, R. Durbin, & Y. S. Song. 2018. Efficiently inferring the demographic history of many populations with allele count data. bioRxiv:1–29. Kamvar, Z. N., J. F. Tabima, & N. J. Grünwald. 2014. Poppr: an R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. PeerJ 2:e281. Karger, D. N., O. Conrad, J. Böhner, T. Kawohl, H. Kreft, R. W. Soria-Auza, N. E. Zimmermann, H. P. Linder, & M. Kessler. 2017. Climatologies at high resolution for the earth’s land surface areas. Scientific Data 4:170122. Klink, C. A., & R. B. Machado. 2005. Conservation of the Brazilian Cerrado. Conservation Biology 19:707–713. Kopelman, N. M., J. Mayzel, M. Jakobsson, N. A. Rosenberg, & I. Mayrose. 2015. Clumpak: a program for identifying clustering modes and packaging population structure inferences across K. Molecular Ecology Resources 15:1179–1191. De La Riva, I., J. Köhler, S. Lötters, & S. Reichle. 2000. Ten years of research on Bolivian amphibians: updated checklist, distribution, taxonomic problems, literature and iconography. Revista Española de Herpetologia 14:19–164. De La Riva, I., R. Marquez, & J. Bosch. 1996. Advertisement calls of four Microhylid frogs from Bolivia (Amphibia, Anura). American Midland Naturalist 136:418–422. Leaché, A. D., B. L. Banbury, J. Felsenstein, A. Nieto-Montes de Oca, & A. Stamatakis. 2015. Short tree, long tree, right tree, wrong tree: new acquisition bias corrections for inferring SNP phylogenies. Systematic Biology 64:1032–1047. Leaché, A. D., M. K. Fujita, V. N. Minin, & R. R. Bouckaert. 2014. Species Delimitation using Genome-Wide SNP Data. Systematic Biology 63:534–542. Ledo, R. M. D., & G. R. Colli. 2017. The historical connections between the Amazon and the Atlantic Forest revisited. Journal of Biogeography 44:2551–2563. Lewis, P. O. 2001. A likelihood approach to estimating phylogeny from discrete morpholoical character data. Systematic Biology 50:913–925. Lutz, B. 1973. Brazilian species of Hyla. University of Texas Press, Austin. Lynch Alfaro, J. W., J. P. Boubli, L. E. Olson, A. Di Fiore, B. Wilson, G. A. Gutiérrez- Espeleta, K. L. Chiou, M. Schulte, S. Neitzel, V. Ross, D. Schwochow, et al. 2012. Explosive Pleistocene range expansion leads to widespread Amazonian sympatry between robust and gracile capuchin monkeys. Journal of Biogeography 39:272–288. Maciel, N. M., R. G. Collevatti, G. R. Colli, & E. F. Schwartz. 2010. Late Miocene diversification and phylogenetic relationships of the huge toads in the Rhinella marina (Linnaeus, 1758) species group (Anura: Bufonidae). Molecular Phylogenetics and Evolution 57:787–797. McCormack, J. E., S. M. Hird, A. J. Zellmer, B. C. Carstens, & R. T. Brumfield. 2013. Applications of next-generation sequencing to phylogeography and phylogenetics. Molecular Phylogenetics and Evolution 66:526–538. Mendes, J. C., & S. Petri. 1971. Geologia do Brasil. In.: Instituto nacional do livro, Rio de Janeiro.

110

Miranda-Ribeiro, A. de. 1920. Os engystomatideos do Museu Paulista (com um gênero e três especies novos). Revista do Museu Paulista, São Paulo:280–288. Miranda, N. E. de O., N. M. Maciel, M. S. Lima-Ribeiro, G. R. Colli, C. F. B. Haddad, & R. G. Collevatti. 2019. Diversification of the widespread neotropical frog Physalaemus cuvieri in response to Neogene-Quaternary geological events and climate dynamics. Molecular Phylogenetics and Evolution 132:67–80. De Moraes-Barros, N., J. A. B. Silva, C. Y. Miyaki, & J. S. Morgante. 2006. Comparative phylogeography of the atlantic forest endemic sloth (Bradypus torquatus) and the widespread three-toed sloth (Bradypus variegatus) (Bradypodidae, Xenarthra). Genetica 126:189–198. Moreira, G., & L. Barreto. 1997. Seasonal variation in nocturnal calling activity of a savanna anuran community in central Brazil. Amphibia-Reptilia 18:49–57. Morrone, J. J. 2009. Evolutionary biogeography: an integrative approach with case studies. In.: Columbia University Press, New York. Morrone, J. J. 2014. Cladistic biogeography of the Neotropical region: identifying the main events in the diversification of the terrestrial biota. Cladistics-the International Journal of the Willi Hennig Society 30:202–214. Motta, P. E. F., N. Curi, & D. P. Franzmeier. 2002. Relation of soils and geomorphic surfaces in the Brazilian cerrado. In: The cerrados of Brazil: ecology and natural history of a neotropical savanna. P. S. Oliveira and R. J. Marquis (eds.). Columbia University Press, New York, pp. 13–32. Myers, N., R. A. Mittermeier, C. G. Mittermeier, G. A. B. da Fonseca, & J. Kent. 2000. Biodiversity hotspots for conservation priorities. Nature 403:853–858. Napoli, M. F., & U. Caramaschi. 1999. Geographic variation of Hyla rubicundula and Hyla anataliasiasi, with a description of a new species (Anura, Hylidae). Alytes 16:165–189. Napoli, M. F., & U. Caramaschi. 2000. Description and variation of a new Brazilian species of the Hyla rubicundula group (Anura:Hyldae). Alytes 17:165–184. Neuwirth, E. 2014. RColorBrewer. R package version1.1-2. Nimer, E. 1989. Climatologia do Brasil. In.: 2nd edition. Fundação Instituto Brasileiro de Geografia e Estatística - IBGE, Rio de Janeiro, Brasil. Nogueira, C., G. R. Colli, & M. Martins. 2009. Local richness and distribution of the lizard fauna in natural habitat mosaics of the Brazilian Cerrado. Austral Ecology 34:83–96. Nogueira, C., S. Ribeiro, G. C. Costa, & G. R. Colli. 2011. Vicariance and endemism in a Neotropical savanna hotspot: distribution patterns of Cerrado squamate reptiles. Journal of Biogeography 38:1907–1922. Oliveira-Filho, A. T., & A. J. Ratter. 2002. Vegetation physiognomies and woody flora of the Cerrado biome. In: The Cerrados of Brazil: Ecology and Natural History of a Neotropical Savanna. P. S. Oliveira and R. J. Marquis (eds.). Columbia University Press, New York, pp. 91–120. Oliveira-Filho, J. C. de, & A. A. Giaretta. 2006. Tadpole and advertisement call of Chiasmocleis albopunctata (Anura, Microhylidae) from Brazil. Zootaxa 1353:63–68.

111

Oliveira, E. F., M. Gehara, V. A. São-Pedro, G. C. Costa, F. T. Burbrink, G. R. Colli, M. T. Rodrigues, & A. A. Garda. 2018. Phylogeography of Muller’s termite frog suggests the vicariant role of the Central Brazilian Plateau. Journal of Biogeography 45:2508–2519. Oliveira, P. S., & R. J. Marquis. 2002. The Cerrados of Brazil: ecology and natural history of a neotropical savanna. In.: Columbia University Press, New York. Otto-bliesner, B. L., B. L. Otto-bliesner, S. J. Marshall, J. T. Overpeck, & G. H. Miller. 2009. Simulating Arctic Climate Warmth and Icefield Retreat in the Last Interglaciation Bette. Science 311:1751–1755. Pattengale, N. D., M. Alipour, O. R. P. Bininda-emonds, B. M. E. Moret, & A. Stamatakis. 2010. How Many Bootstrap Replicates Are Necessary ?. Journal of Computational Biology 17:337–354. Patterson, N., P. Moorjani, Y. Luo, S. Mallick, N. Rohland, Y. Zhan, T. Genschoreck, T. Webster, & D. Reich. 2012. Ancient admixture in human history. Genetics 192:1065– 1093. Pearson, P. N., & M. R. Palmer. 2000. Nature Atmospheric carbon dioxide concentrations over the past 60 million years. Nature 406:695–699. Pearson, R. G. 2008. Species’ distribution modeling for conservation educators and practitioners. Synthesis. American Museum of Natural History. Available at http://ncep.amnh.org. Peloso, P. L. V, M. J. Sturaro, M. C. Forlani, P. Gaucher, A. P. Motta, & W. C. Wheeler. 2014. Phylogeny, taxonomic revision, and character evolution of the genera Chiasmocleis and Syncope (Anura, Microhylidae) in Amazonia, with descriptions of three new species. Bulletin of the American Museum of Natural History:1–112. Peterson, B. K., J. N. Weber, E. H. Kay, H. S. Fisher, & H. E. Hoekstra. 2012. Double digest RADseq: an inexpensive method for De Novo SNP discovery and genotyping in model and non-model species. PLoS ONE 7:e37135. Peterson, T., & D. A. Vieglais. 2001. Predicting species invasions using ecological niche modeling: new approaches from bioinformatics attack a pressing problem. Bioscience 51:363. Phillips, S. J., R. P. Anderson, & R. E. Schapire. 2006. Maximum entropy modeling of species geographic distributions. Ecological Modelling 190:231–259. Prado, C. P., C. F. B. Haddad, & K. R. Zamudio. 2012. Cryptic lineages and Pleistocene population expansion in a Brazilian Cerrado frog. Molecular Ecology 21:921–941. Prates, I., M. J. Hickerson, A. T. Xue, M. T. Rodrigues, D. F. Alvarado-Serrano, A. C. Carnaval, & J. L. Brown. 2016a. Inferring responses to climate dynamics from historical demography in neotropical forest lizards. Proceedings of the National Academy of Sciences USA 113:7978–7985. Prates, I., D. Rivera, M. T. Rodrigues, & A. C. Carnaval. 2016b. A mid-Pleistocene rainforest corridor enabled synchronous invasions of the Atlantic Forest by Amazonian anole lizards. Molecular Ecology 25:5174–5186. Pritchard, J. K., M. Stephens, & P. Donnelly. 2000. Inference of population structure using

112

multilocus genotype data. Genetics 155:945–959. QGIS Development Team. 2015. QGIS Geographic Information System. Open source geospatial foundation project. R Development Core Team. 2017. R: a language and environment for statistical computing. [Online.] Available at http://www.r-project.org. Ratter, J. A., J. F. Ribeiro, & S. Bridgewater. 1997. The Brazilian cerrado vegetation and threats to its biodiversity. Annals of Botany 80:223–230. Reinhardt, J. T., & C. F. Lütken. 1862. Bidrag til Kundskab om Brasiliens Padder og Krybdyr. Förste Afdeling: Padderne og Öglerne. Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening i Kjøbenhavn 2:143–242. Reynolds, J., B. S. Weir, & C. C. Cockerham. 1983. Estimation of the coancestry coefficient: Basis for a short-term genetic distance. Genetics 105:767–779. Rodrigues, M. T. 2005. A biodiversidade dos Cerrados: conhecimento atual e perspectivas, com uma hipótese sobre o papel das matas galerias na troca faunística durante ciclos climáticos. In: Cerrado: ecologia, biodiversidade e conservação. A. Scariot, J. C. Sousa-Silva and J. M. Felfili (eds.). Ministério do Meio Ambiente, Brasília, Distrito Federal, pp. 235–246. Rull, V. 2008. Speciation timing and neotropical biodiversity: the Tertiary-Quaternary debate in the light of molecular phylogenetic evidence. Molecular Ecology 17:2722–2729. Rylands, A. B., & K. Brandon. 2005. Brazilian protected areas. Conservation Biology 19:612– 618. de Sá, R. O., J. W. Streicher, R. Sekonyela, M. C. Forlani, S. P. Loader, E. Greenbaum, S. Richards, & C. F. B. Haddad. 2012. Molecular phylogeny of microhylid frogs (Anura: Microhylidae) with emphasis on relationships among New World genera. BMC Evolutionary Biology 12:241. de Sá, R. O., J. F. R. Tonini, H. van Huss, A. Long, T. Cuddy, M. C. Forlani, P. L. V. Peloso, H. Zaher, & C. F. B. Haddad. 2019. Multiple connections between Amazonia and Atlantic Forest shaped the phylogenetic and morphological diversity of Chiasmocleis Mehely, 1904 (Anura: Microhylidae: Gastrophryninae). Molecular Phylogenetics and Evolution 130:198–210. Salgado, G. S. M., & M. Galinkin. 2008. Programa o homem e a biosfera da UNESCO. Águas Emendadas. Sánchez, J. M., G. P. Pesci, S. M. Leão, & N. Pelegrin. 2014. First record of the genus Chiasmocleis Mèhelÿ, 1904 (Anura, Microhylidae) in Argentina. Check List 10:962–964. Santos, J. C., L. A. Coloma, K. Summers, J. P. Caldwell, R. Ree, & D. C. Cannatella. 2009. Amazonian amphibian diversity is primarily derived from Late Miocene Andean lineages. PLoS Biology 7:448–461. Santos, M. G., C. Nogueira, L. G. Giugliano, & G. R. Colli. 2014. Landscape evolution and phylogeography of Micrablepharus atticolus (Squamata, ), an endemic lizard of the Brazilian Cerrado. Journal of Biogeography 41:1506–1519. Santos, T. G. dos, D. de C. Rossa-Feres, & L. Casatti. 2007. Diversidade e distribuição espaço-temporal de anuros em região com pronunciada estação seca no sudeste do Brasil.

113

Iheringia 97:37–49. Silva, F. R. da, V. H. M. do Prado, T. da S. Vasconcelos, T. G. dos Santos, & D. de C. Rossa-Feres. 2009. Amphibia, Anura, Microhylidae, Chiasmocleis albopunctata: filling gap and geographic distribution map. Check List 5:314–316. Silva, G. R. da, J. B. Vieira, & M. R. Alves. 2011. Amphibia, Anura, Hylidae, Dendropsophus rubicundulus (Reinhardt and Lütken, 1862): distribution extension in northeastern Brazil. Check List 7:843–844. Silva, J. M. C. 1997. Endemic bird species and conservation in the Cerrado Region, South America. Biodiversity and Conservation 6:435–450. Smith, P., P. Cacciali, K. Atkinson, H. Pheasey, & M. Motte. 2012. New distributional records of amphibians for Departamento San Pedro, Paraguay (Amphibia). Check List 8:903–907. Solomon, S. E., M. Bacci, J. Martins, G. G. Vinha, & U. G. Mueller. 2008. Paleodistributions and comparative molecular phylogeography of leafcutter ants (Atta spp.) provide new insight into the origins of Amazonian diversity. PLoS ONE 3:e2738. Stamatakis, A. 2014. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30:1312–1313. Swofford, D. L. 2002. PAUP*. Phylogenetic Analyses Using Parsimony (*and other methods). Version 4. Sunderland, Sinauer Associates. Urbanek, S. 2015. rJava: Low-level R to Java interface. R package version 0.9-6. Valdujo, P. H., D. L. Silvano, G. Colli, & M. Martins. 2012. Anuran Species Composition and Distribution Patterns in Brazilian Cerrado, a Neotropical Hotspot. South American Journal of Herpetology 7:63–78. Vallinoto, M., F. Sequeira, D. Sodre, J. A. R. Bernardi, I. Sampaio, & H. Schneider. 2010. Phylogeny and biogeography of the Rhinella marina species complex (Amphibia, Bufonidae) revisited: implications for Neotropical diversification hypotheses. Zoologica Scripta 39:128–140. Vanzolini, P. E., & E. E. Williams. 1981. The vanishing refuge: a mechanism for ecogeographic speciation. Papeis Avulsos De Zoologia 34:251–255. Vasconcellos, M. M., G. R. Colli, J. N. Weber, E. M. Ortiz, M. T. Rodrigues, & D. C. Cannatella. 2019. Isolation by instability: Historical climate change shapes population structure and genomic divergence of treefrogs in the Neotropical Cerrado savanna. Molecular Ecology. Accepted Author Manuscript. Vasconcelos, T. D. S., T. G. Dos Santos, D. D. C. Rossa-Feres, & C. F. B. Hadadd. 2011. Spatio-temporal distribution of calling male frogs in Morro Do Diabo State Park (Southeastern Brazil): implications for conservation and management in a Seasonally Dry Tropical Forest. South American Journal of Herpetology 6:107–118. Vaz-Silva, W., & N. M. Maciel. 2011. A new cryptic species of Ameerega (Anura: Dendrobatidae) from Brazilian Cerrado. Zootaxa:57–68. Vitt, L. J. 1991. An introduction to the ecology of Cerrado lizards. Journal of Herpetology 25:79–90.

114

Werneck, F. D. P. 2011. The diversification of eastern South American open vegetation biomes: Historical biogeography and perspectives. Quaternary Science Reviews 30:1630–1648. Werneck, F. P., T. Gamble, G. R. Colli, M. T. Rodrigues, & J. W. Sites. 2012a. Deep diversification and long-term persistence in the South American “dry diagonal”: integrating continent-wide phylogeography and distribution modeling of geckos. Evolution; international journal of organic evolution 66:3014–3034. Werneck, F. P., L. G. Giugliano, R. G. Collevatti, & G. R. Colli. 2009. Phylogeny, biogeography and evolution of clutch size in South American lizards of the genus Kentropyx (Squamata: Teiidae). Molecular Ecology 18:262–278. Werneck, F. P., C. Nogueira, G. R. Colli, J. W. Sites, & G. C. Costa. 2012b. Climatic stability in the Brazilian Cerrado: implications for biogeographical connections of South American savannas, species richness and conservation in a biodiversity hotspot. Journal of Biogeography 39:1695–1706. Wiens, J. J., & C. H. Graham. 2005. Niche conservatism: integrating evolution, ecology, and conservation biology. Annual Review of Ecology Evolution and Systematics 36:519–539. Wright, S. 1943. Isolation by Distance. Genetics 28:114 LP – 138. Zamudio, K. R., R. C. Bell, & N. A. Mason. 2016. Phenotypes in phylogeography: Species’ traits, environmental variation, and vertebrate diversification. Proceedings of the National Academy of Sciences USA 113:8041–8048.

115

VITA

PROFESSIONAL PREPARATION

University of Mississippi, United States:

2014 - 2019 Ph.D. in Biology. Dissertation Title: “Looking into the past: the diversification of amphibians in the Neotropical savannas.” Dissertation Advisor: PhD. Brice Noonan. 2015 - 2019 Interdisciplinary Certificate in Applied Statistics (ICAS)

Universidade de Brasília, Brazil:

2013 Environmental GIS, Certificate in Applied Geographic Information Systems (GIS) 2010 M.S. Ecology 2008 Biology, Bachelor of Science and Bachelor of Education

PROFESSIONAL APPOINTMENTS

06/2011 - 12/2013 Collections manager, Herpetological collection at Universidade de Brasília (CHUNB), Brazil. 06/2010 - 05/2011 Laboratory technician, Laboratory of Plant Anatomy at Universidade de Brasília, Brazil. 04/2009 - 05/2010 and 11/2010 - 02/2011 Instructional Professor for an online biology degree program (Licenciatura em biologia à distância – LicBio) at Universidade de Brasília, Brazil.

PUBLICATIONS

Domingos F. M. C. B, Arantes Í. C., Bosque R. J., Santos M.G. 2017. Nesting in the lizard Phyllopezus pollicaris (Squamata: Phyllodactylidae) and a phylogenetic perspective on communal nesting in the family. Phyllomedusa 16:255–267. Arantes, Í. C., Vasconcellos, M. M., Boas, T. C. V., Veludo, L. B. A. & Colli, G. R. 2015. Sexual Dimorphism, Growth, and Longevity of Two Toad Species (Anura, Bufonidae) in a Neotropical Savanna. Copeia. 103:329-342. Domingos, F. M. C. B., Arantes Í. C., Cavalcanti, D. R., Jotta, P. A. C. V. 2014 Shelter from the sand: Microhabitat Selection by the Bromelicolous Tree Frog cuspidatus (Anura, Hylidae) in a Brazilian Restinga. North-Western Journal of Zoology 11 (1): 27- 33.

116

Manuscripts in Preparation

Arantes, Í. C., Zimbres, B. Q., Aquino, P. P. U., Colli, G. R.; Toledo, L. F. Amphibian Chytrid Fungus: a Conservation Concern in Brazil.

AWARDS AND HONORS

2017 Poster “Phylogeography of Neotropical savanna frogs” at Neuroscience Research Showcase, University of Mississippi, University, MS, (Honorable mention). 2016 Poster “Phylogeography of Neotropical savanna frogs” at Evolution Meeting at Austin, TX, July. Support Society for The Study of Evolution, SSE, ($500 travel stipend award). 2013 Poster “Dimorfismo Sexual, Crecimiento y Longevidad de dos Especies del género Rhinella (Anura, Bufonidae) en el Cerrado” at XIV Congreso Argentino de Herpetología, Puerto Madryn, Argentina, September. Support Research Foundation of Federal District (Fundação de Apoio a Pesquisa do Distrito Federal – FAPDF), (R$ 7,750.00). 2011 Poster “Sexual dimorphism and chromatic polymorphism in Hoplocercus spinosus (Iguania, Hoplocercidae)” at IX Latin American Congress of Herpetology (CLAH), July. Support Research Foundation of Federal District (Fundação de Apoio a Pesquisa do Distrito Federal – FAPDF), (R$ 2,000.00).

GRANTS AND FELLOWSHIPS

2017 Conference travel grant ($600), University of Mississippi, Department of Biology and Graduate school. 2016 Conference travel grant ($600), University of Mississippi, Department of Biology and Graduate school. 2014 Conference travel grant ($600), University of Mississippi, Department of Biology and Graduate school.

M.S. Fellowship

Conselho Nacional de Desenvolvimento Científico e Tecnológico, CNPq, Brazil (R$ 28,800.00).

PhD. scholarship

Coordenação de Aperfeiçoamento de Pessoal de Nível Superior, CAPES, Brazil. This pays for fieldwork, tuition and fees, airplane tickets and $1300.00 monthly stipend.

CONFERENCE ACTIVITIES

117

Conferences/ Symposia organized

2017 XI Latin American Congress of Herpetology, symposium coordinator of “Evolutionary and ecological processes affecting Neotropical frog communities”, Quito, , July 24-28. 2009 IV Brazilian Herpetological Meeting, Pirenópolis, GO, Brazil, July 12-17. 2005 XIX Annual Meeting of the Society for Conservation Biology, Brasília, DF, Brazil, July 15-19.

Talk

2017 “Comparative Phylogeography of Neotropical savanna endemic frogs (Chiasmocleis albopunctata, Dendropsophus rubicundulus and Physalaemus nattereri)” at XI Latin American Congress of Herpetology, Quito, Ecuador, July 24-28.

Poster presented

2017 “Phylogeography of Neotropical savanna frogs” at Neuroscience Research Showcase, University of Mississippi, University, MS, March 31. 2016 “Phylogeography of Neotropical savanna frogs” at Evolution 2016 in Austin, TX. Support: Society for The Study of Evolution, SSE, June 17-21. 2014 “Is the susceptibility to chytridiomycosis related to niche conservatism?” at Evolution 2014 in Raleigh, NC. Support: Society for The Study of Evolution, SSE, June 20-24. 2013 “Dimorfismo Sexual, Crecimiento y Longevidad de dos Especies del género Rhinella (Anura, Bufonidae) en el Cerrado” at XIV Congreso Argentino de Herpetología, Puerto Madryn, Argentina, September 17-20. 2011 “Sexual dimorphism and chromatic polymorphism in Hoplocercus spinosus (Iguania, Hoplocercidae)” at IX Latin American Congress of Herpetology (CLAH), July 17-22.

TEACHING EXPERIENCE

2018 Teaching Assistant, Genetics, University of Mississippi. 2018 Teaching Assistant, Inquiry into Life Laboratory I, University of Mississippi. 2017 Teaching Assistant, Biological sciences 1 – Laboratory, University of Mississippi. 2016 Teaching Assistant, Inquiry into Life Laboratory I, University of Mississippi. 2016 Teaching Assistant, Biological sciences 1 – Laboratory, University of Mississippi. 2016 Teaching Assistant, Genetics, University of Mississippi. 2014 Teaching Assistant, Inquiry into Life Laboratory I, University of Mississippi. 2008 Teaching Assistant, Biology of Reptiles, Universidade de Brasília, Brazil. 2009 Teaching Assistant, Statistics Applied to Ecology, Universidade de Brasília, Brazil. 2007 Teaching Assistant, Vertebrate Zoology, Universidade de Brasília, Brazil.

118

RESEARCH EXPERIENCE

Ph.D. research

My work seeks to unravel the evolutionary processes that led to diversification of the amphibians of the Neotropical savannas of Brazil. Evolutionary patterns are being investigated by comparing the phylogeography and biogeography of endemic and widely distributed amphibian species using next generation sequencing.

Master’s research

I studied the sexual dimorphism, growth, and longevity of Rhinella schneideri and R. rubescens in a pond in the Neotropical savannas of central Brazil. The study was based on 3 years of weekly sampling at the study pond to measure the growth rate of all individuals for both species. These data were combined with morphological data of sexual dimorphism collected in the herpetological collection at Universidade de Brasília, Brazil.

SERVICE TO PROFESSION

Manuscript reviewer 2018 Journal of Natural History 2016 Herpetological Conservation and Biology 2016 Russian Journal of Herpetology

COMMUNITY INVOLVEMENT AND OUTREACH

2009 “Unraveling the secret of the snakes” at IV Brazilian Herpetological Meeting, Pirenópolis, GO, Brazil. The main goal of the presentation was to promote environmental education for the community, helping them to identify venomous snakes and prevent bites. 2010 “Six years studying frogs in Brasília, Brazil” at a local environmental magazine whose main goal is to promote environmental education for the community that lives near the protected area. The article discussed the results of long-term research at the Ecological station of Águas Emendadas, Brasília, Brazil. Arantes, Í. C.; Vilas Boas, T. C.; Colli, G. R. Seis anos estudando os sapos do DF. Educação Ambiental, Brasília - DF, p. 30 - 31, 01 ago. 2010.

RELATED PROFESSIONAL SKILL

Data analysis: advanced skills using R environment, Geographic Information Systems

119

using R, QGIS and ArcGIS, species distribution modeling using Maxent and R, bioinformatics using High Performance Computers.

Laboratory: morphological measurements, next generation sequencing, molecular laboratory techniques (e.g., DNA extraction, PCR, genotyping).

Fieldwork: identification, capture and euthanasia of herpetological specimens, collection tissue samples and specimen preservation to deposit in the museums.

LANGUAGES

Portuguese – native speaker (read, listen, speak and write). English – fluent (read, listen, speak and write). Spanish – good (read, listen, speak and write).

PROFESSIONAL MEMBERSHIP AND AFFILIATIONS

2012 Sociedade Brasileira de Herpetologia (SBH) 2014 Society for the Study of Evolution (SSE).

120