11.1.125-Magalhaes Et Al. 2022

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Molecular Phylogenetics and Evolution 169 (2022) 107398

Contents lists available at ScienceDirect

Molecular Phylogenetics and Evolution


journal homepage: www.elsevier.com/locate/ympev

Ecological divergence and synchronous Pleistocene diversification in the


widespread South American butter frog complex
Felipe de M. Magalhães a, b, *, Felipe Camurugi c, Mariana L. Lyra d, Diego Baldo e,
Marcelo Gehara b, Célio F.B. Haddad d, Adrian A. Garda f
a
Programa de Pós-Graduação em Ciências Biológicas, Universidade Federal da Paraíba–UFPB, Centro de Ciências Exatas e da Natureza, Cidade Universitária, 58000-
000 João Pessoa, Paraiba, Brazil
b
Earth and Environmental Sciences, Ecology and Evolution, Rutgers University-Newark 195 University Ave, Newark, NJ 07102, USA
c
Instituto de Biociências, Universidade Federal de Mato Grosso do Sul, Cidade Universitária, 79070-900, Campo Grande, Mato Grosso do Sul, Brazil
d
Instituto de Biociências, Universidade Estadual Paulista, Campus Rio Claro, Departamento de Biodiversidade e Centro de Aquicultura (CAUNESP), Laboratório de
Herpetologia, Cx. Postal 199, 13506-900 Rio Claro, São Paulo, Brazil
e
Instituto de Biología Subtropical (IBS, CONICET-UNaM), Laboratorio de Genética Evolutiva, Facultad de Ciencias Exactas, Universidad Nacional de Misiones, Félix de
Azara 1552, CPA N3300LQF Posadas, Misiones, Argentina
f
Laboratório de Anfíbios e Répteis (LAR), Departamento de Botânica e Zoologia da Universidade Federal do Rio Grande do Norte, Campus Universitário. Lagoa Nova,
59078-900 Natal, Rio Grande do Norte, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Phylogeographic studies primarily focus on the major role of landscape topography in driving lineage diversi­
Anura fication. However, populational phylogeographic breaks may also occur as a result of either niche conservatism
Atlantic Forest or divergence, in the absence of geographic barriers to gene flow. Furthermore, these two factors are not
Cryptic diversity
mutually exclusive and can act in concert, making it challenging to evaluate their relative importances on
Divergence Times
explaining genetic variation in nature. Herein, we use sequences of two mitochondrial and four nuclear genes to
Niche divergence
Niche modeling investigate the timing and diversification patterns of species pertaining to the Leptodactylus latrans complex,
Species tree which harbors four morphologically cryptic species with broad distributions across environmental gradients in
Species delimitation eastern South America. The origin of this species complex dates back to the late Miocene (ca. 5.5 Mya), but most
diversification events occurred synchronically during the late Pleistocene likely as the result of ecological
divergence driven by Quaternary climatic oscillations. Further, significant patterns of environmental niche di­
vergences among species in the L. latrans complex imply that ecological isolation is the primary mode of genetic
diversification, mostly because phylogenetic breaks are associated with environmental transitions rather than
topographic barriers at both species and populational scales. We provided new insights about diversification
patterns and processes within a species complex of broadly and continuously distributed group of frogs along
South America.

1. Introduction now evoking other isolation mechanisms that are not topography-driven
and that are also important for promoting genetic differentiation (Irwin,
Understanding patterns and processes underlying the diversification 2002). For instance, environmental and ecological divergences between
of organisms is of utmost interest to evolutionary biologists, which has lineages can also lead to phylogeographic structure posed by natural
intensified with the growing field of phylogeography. Since their dawn selection on populations in ecologically divergent habitats (Wang and
phylogeographic studies have primarily focused on the spatial distri­ Summers, 2010; Pyron and Burbrink, 2010; Wang and Bradburd, 2014).
bution of lineages on a landscape and associated geographic factors such Likewise, topographic heterogeneity (Rodríguez et al., 2015), climatic
as geographic distance, topography (e.g., geological events), and phys­ oscillations (Byrne, 2008), sea-level fluctuations (Fazolato et al., 2017;
ical barriers to genetic differentiation throughout lineages evolutionary Senczuk et al., 2019) and biotic interactions (Waters, 2011) have also
history (Avise, 2000). Nevertheless, recent phylogeographic studies are been increasingly reported as a primary mode of genetic diversification

* Corresponding author at: Earth and Environmental Sciences, Ecology and Evolution, Rutgers University-Newark 195 University Ave, Newark, NJ 07102, USA.
E-mail address: felipemm17@gmail.com (F.M. Magalhães).

https://doi.org/10.1016/j.ympev.2022.107398
Received 11 August 2021; Received in revised form 31 October 2021; Accepted 15 November 2021
Available online 11 January 2022
1055-7903/© 2022 Elsevier Inc. All rights reserved.
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

worldwide. Although alternative drivers of genetic differentiation are reviewed by Magalhães et al., 2020). These four species cluster into a
becoming more widely appreciated (Gehara et al., 2017a; Vasconcellos highly supported monophyletic clade (e.g., share a most recent common
et al., 2019; Camurugi et al., 2021), relatively few studies have exam­ ancestor) within the L. latrans species group. They also exhibit
ined the roles of geographic and ecological factors in explaining phy­ remarkably similar external morphologies and advertisement calls
logeographic patterns at different stages of diversification and (except for L. payaya), which are of paramount importance for intra­
throughout heterogeneous habitats at continental scales (e.g., Pyron and specific recognition in frogs, acting as isolating mechanisms precluding
Burbrink, 2009; Wang et al., 2013; Myers et al., 2020). secondary contact and restricting hybridization (Wells, 2007). Never­
The South American Neotropical region is one of the most species- theless, multilocus molecular data revealed deep genetic structure, large
rich regions of the world, harboring several hotspots of biodiversity genetic distances and apparent low gene exchange (Magalhães et al.,
(Mittermeier et al., 1998) and highly diverse in habitats (e.g., Ab’Saber, 2020). Two species are endemic to the AF (L. latrans and L. paranaru),
1977; Vieira et al., 2015). Diversification processes along this region are while L. luctator partially occurs across higher elevation sites within the
mostly linked to geomorphological features (such as rivers and barriers) AF, but also at lower elevation of open formations within Pampas, Chaco
and climatic fluctuations (Hoorn et al., 2010; Rull, 2011; Turchetto- Monte, and Pantanal biomes. The fourth species, L. payaya, is mainly
Zolet et al., 2013; Smith et al., 2014). More recently, Carnaval and distributed throughout open formations associated to the Chapada
Moritz (2008) proposed that during glacial periods of the Pleistocene Diamantina ecoregion and vicinities (for a characterization of this
areas of climatic stability along the Atlantic Forest (AF) acted as refuge ecoregion see Lima and Nolasco, 2015) along interior of Bahia State,
in a fragmented landscape as forests retracted due to colder and drier northeastern Brazil, which is surrounded by the Caatinga, a seasonally
climatic conditions, leading to population isolation and genetic diver­ dry forest biome (Lima and Nolasco, 2015). Despite being restricted to
sification, hereafter referred as Carnaval-Moritz model (CM model). the AF, L. latrans and L. paranaru are not forest-dependent species, and
Therefore, AF species underwent a common set of processes resulting all four species share similar reproductive patterns: they all reproduce in
from fragmentation within glacial refugia (Carnaval et al., 2009; Tonini temporary ponds in open fields or forest edges, are abundant and
et al., 2013), range expansions via postglacial colonization (e.g., Mar­ generalist feeders, preying mostly on invertebrates and also vertebrates
tins, 2011; Firkowski et al., 2016), and are likely experiencing secondary such as small to medium-sized frog species. Moreover, all four species
contacts among historically divergent lineages (Fitzpatrick et al., 2009; are mostly low elevation specialists (predominantly occurring in areas
Sabbag et al., 2018). Indeed, the CM model expectations have been below 700 m above sea level), while L. luctator can reach high plain sites
recurrently corroborated in papers focusing on patterns of genetic di­ at 1400 m above sea level across Brazil mainland or 1600 m in the Chaco
versity among distinct vertebrates and invertebrates organisms distrib­ of Argentina (Magalhães et al., 2020).
uted along the AF (Batalha-Filho et al., 2010; D’Horta et al., 2011; The species in the L. latrans complex are widely and continuously
Carnaval et al., 2014) and raised as a possible explanation for the high distributed across environmental (e.g., from seasonally dry to tropical
genetic diversity and elevated species richness found along the AF when moist forests) and altitudinal (from sea-level to highlands) gradients,
compared to other biomes (Morellato and Haddad, 2000; Carnaval and show strong signs of geographic and genetic structure, and low
Moritz, 2008). Conversely to that expectations, Leite et al. (2016) pro­ morphologic and acoustic divergences, making them an interesting
posed that during glacial cycles the AF expanded towards the exposed biological group to address how geographic and ecological isolation
continental shelves as global sea-levels decreased after the Last Inter­ shaped evolutionary history within this group of frogs. With recent ad­
glacial (LIG) period, favoring AF associated taxa to expand instead of vances in methods for assessing niche divergence using ecological niche
retract to climatically stable forest fragments (as predicted by CM models (ENMs) allied to coalescent genetic analyses (Carstens and
model), which played a minor role in shaping genetic diversification Richards, 2007), it is now possible to weigh the relative importance of
along the AF. these two factors in explaining patterns of phylogeographic structure
The diversification processes linked to Pleistocene climate oscilla­ among groups (Thorpe et al., 2008; Wang et al., 2013; Maia-Carvalho
tions are still controversial (Rull, 2008, 2011) and may not explain ge­ et al., 2018).
netic diversity across AF organisms (Thomé et al., 2014). For instance, In this paper, we investigate population genetic structure, divergence
sea-level transgressions/regressions cycles intensified during repeated times and the effects of late Pleistocene climatic changes on the de­
Pleistocene glaciation cycles (Hansen et al., 2013), which certainly mographic history and genetic diversity in a complex of cryptic species
caused modifications to the coastline (Suguio et al., 1985), and could broadly distributed along the eastern region of South America. Our goals
have promoted lineage diversification by creating physical barriers are to clarify geographic patterns of genetic diversity and the processes
along lowland continuously distributed taxa (e.g., Fitzpatrick et al., involved in the diversification of Leptodactylus latrans species complex.
2009; Fazolato et al., 2017; Gehara et al., 2017a). The presence of major We used a multi-locus dataset (two mitochondrial and four nuclear loci)
rivers along eastern South America have also been pointed as an to characterize geographic population structure relative to the presence
important mechanism for allopatric diversification of several taxa acting of putative barriers (geographic) and transitional environmental areas
as a hard or permeable barrier to gene flow (Bruschi et al., 2019; Sotelo- (ecologic) in the range of these frog species. If diversification is driven by
Muñoz et al., 2020; Bocalini et al., 2021). Hence, geographic and allopatric isolation, we expect that both geographic and temporal phy­
ecological isolating mechanisms can act in concert, precluding the logeographic breaks match major topographic barriers to gene flow.
recognition of the primary mode of lineage diversification. This becomes Instead, if ecological isolation mechanisms are playing a major role in
especially intricate in regions where topography, climatic oscillations shaping diversification we expect phylogeographic breaks to be coinci­
and biomes are changing constantly and concomitantly, as is the case of dent with environmental transition areas and that environmental niches
the South America Neotropical region (Vuilleumier, 1971; Hoorn et al., are also divergent between sister lineages.
2010). Species that are distributed widely across environmentally and
geographically heterogeneous landscapes offer opportunities to inves­ 2. Material and methods
tigate distinct patterns of diversification, and are particularly valuable
for understanding the role of geographic and ecological factors on ge­ 2.1. Sampling and sequencing
netic differentiation.
The Leptodactylus latrans complex (also known as butter frogs in Magalhães et al. (2020) extensively sampled and delimited the major
Brazil) encompasses four large-sized (SVL varying from 80 to 120 mm) evolutionary lineages within the Leptodactylus latrans species group,
morphologically cryptic species broadly and continuously distributed which encompasses the L. latrans species complex. To further investigate
throughout 3,000 km across eastern South America with occurrence in genetic diversity and diversification patterns of the four species in the
five distinct biomes (taxonomy and species limits were thoroughly L. latrans complex, we analyzed two mitochondrial (mtDNA) and four

2
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

nuclear (nuDNA) genes fragments for 144 specimens from 131 localities subsequent assignment tests, including individuals with at least two
along eastern South American region (see Table S1 for the list of samples nuclear loci. Moreover, we ran the following assignment tests separately
and localities), representing a subset of the 327 sequences of cytochrome for each of the four species to increase detectability of genetic structure
c oxidase I (COI) dataset provided by Magalhães et al. (2020). Because rather than species limits, which has been shown to be large by mo­
species in this complex present both restricted (e.g., L. payaya and lecular means (Magalhães et al., 2020).
L. paranaru) and wide (e.g., L. latrans and L. luctator) geographic ranges, To assign individuals to genetic clusters, we used a model-based
we have distinct genetic sample sets per species, but we equally and clustering method implemented in STRUCTURE v2.3.4 (Pritchard
continuously sampled across their entire distribution range. Specimens et al., 2000). Using multi-locus genotypic data, STRUCTURE divides
and tissues were obtained from distinct scientific collections from individuals into genetic clusters (K) that both minimizes deviations from
Argentina, Brazil, Paraguay, and Uruguay, which are also listed in Hardy–Weinberg and linkage equilibrium within each cluster and also
Table S1. calculates fractional membership of individuals to each cluster (Q). To
We extracted DNA from muscle/liver tissues using the standard salt convert sequences to STRUCTURE input, we used the program
precipitation methods (Maniatis et al., 1982; Bruford et al., 1992). Po­ xmfa2struct (available at: http://www.xavierdidelot. xtreemhost.com/
lymerase chain reaction (PCR) amplifications were carried out using Taq clonalframe.htm), which encodes each variable site as an allele. We
DNA Polymerase Master Mix (Ampliqon S/A, Denmark) to amplify performed ten independent runs for each K (ranging from 1 to 6 pop­
fragment sequences of the mitochondrial cytochrome c oxidase I (COI) ulations for each species). We ran the analyses for 50,000 iterations as
and 16S rRNA (16S) genes, and nuclear genes fragments of cellular burn-in and 100,000 additional iterations implementing the linkage
myelocytomatosis oncogene exon 2 (cmyc2), tyrosinase precursor (Tyr), model and uncorrelated allele frequencies. The most likely K was
β-Fibrinogen Intron 7 (Fib7), and proopiomelanocortin (POMC) (for determined based on the method of Evanno et al. (2005) via the on-line
primers and PCR protocols see Table S2, Supplementary material). On program Structure Harvester v.0.6.94 (Earl and vonHoldt, 2012).
most occasions, we conducted PCR product cleaning using enzymatic Given the impossibility of validating K = 1 by means of Evanno’s
purifications (shrimp alkaline phosphatase and exonuclease I; Werle methods (Janes et al., 2017), we also estimated the optimal number of
et al., 1994). Purified or unpurified PCR products were sent to Macrogen populations with the Bayesian program Geneland v4.0.8 (Geneland R
Inc. (South Korea) for purification (when needed) and sequencing. We library; Guillot et al., 2005) in the R 3.6.3 environment (R Core Team,
aligned all sequences using MAFFT algorithm (Katoh et al., 2002) with 2018) to complement STRUCTURE assignment analysis. Unlike
default configuration implemented in Geneious v9.1.8 (Kearse et al., STRUCTURE, Geneland also considers geographic location on individual
2012). We sequenced the same genes for individuals of L. viridis, assignment to a given cluster K. We also ran Geneland population
L. bolivianus and L. guianensis to use as outgroups when needed. All new assignment models separately for each of the four species as follows: We
sequences obtained in this study are available at GenBank (Accession performed ten independent runs with K values ranging from 1 to 6
numbers: OM307608–32 and OM338650–939; Table S1, Supplementary populations, assuming an uncorrelated allele frequency, and a spatial
material). model with uncertainty on coordinates. Each run consisted of 5,000,000
We assumed heterozygosity for nuclear sequences when chromato­ MCMC iterations, thinning interval of 5,000 and a burn-in phase of 200
grams contained strong equal double peaks, typically>50% of neigh­ iterations. We assessed the MCMC posterior probability (pp.) plot to
boring homozygotic peaks height. We recovered phase information from determine run convergence and compared it with the posterior estimates
nuclear sequences using PHASE algorithm (Stephens et al., 2001) of the number of populations.
implemented in DNAsp v5 (Librado and Rozas, 2009). Only samples We also implemented the Bayesian Phylogenetics and Phylogeog­
with probability of pairs of alleles in heterozygosity higher than 70% raphy species delimitation method (BPP4) v4.1.3 (Flouri et al., 2018) to
were considered in the following analyses. To estimate the best substi­ validate the genetic breaks identified by both STRUCTURE and Gene­
tution model of each gene fragment (Table S3), we used the Bayesian land analyses combined. BPP is a multi-locus method that implements a
Information Criterion (BIC) implemented in jModelTest 2.1.4 (Darriba reversible jump MCMC algorithm under the multispecies coalescent
et al., 2012). model to delimit species and estimate a species-tree. BPP accommodates
differences in loci mutation rates and effective population sizes related
2.2. Haplotype network genealogies to mitochondrial and nuclear genomes, although analyzing them sepa­
rately is recommended (Flouri et al., 2018). Therefore, we used mito­
To investigate genealogical relationships, we estimated gene trees chondrial loci and three nuclear genes (POMC not included) to test all
using Bayesian inference in BEAST software v2.6.6 (Bouckaert et al., possible clustering hypotheses, but also running BPP with nuclear and
2019). We estimated independent gene trees for each nuclear locus, mitochondrial loci separately. Although we have two mitochondrial
while for the mitochondrial genes (16S + COI) we recovered a single genes, these were concatenated and analyzed as a single locus. We ran
topology by concatenating tree priors in BEAST, but partitioning among BPP with four different prior combinations for population size (theta =
genes in order to accommodate evolutionary rate heterogeneity among θ) and divergence time (tau = τ) parameters as proposed by Leaché and
partitions. We used the most appropriate substitution model for each Fujita (2010): (i) large ancestral population sizes and ancient divergence
gene (Table S3) and ran 20,000,000 generations sampled every 2,000 [IG(3,0.2) for both parameters]; (ii) large population sizes and recent
generations for all five genes using the Coalescent model as tree prior. divergences [IG (3,0.2) for theta; and IG(3,0.002) for tau]; (iii) small
We visually assessed convergence of the Monte Carlo Markov Chain population sizes and recent divergences [IG(3,0.002) for both parame­
(MCMC) runs and effective sample sizes (ESS values > 200) using Tracer ters]; and (iv) small population sizes and deep divergences [IG(3,0.002)
1.7 (Rambaut et al., 2018). We discarded 25% of generations as burn-in, for theta; and IG(3,0.2) for tau]. We ran all prior combinations at least
and the maximum clade credibility tree for each locus was inferred with twice using both algorithms available in BPP (0 and 1) and adjusted the
TreeAnnotator (BEAST package). We used these gene trees (without finetuning to ensure swapping rates between 0.30 and 0.70. We per­
outgroups) to estimate haplotype networks in HAPLOVIEWER formed the A11 analysis, which jointly delimit species and estimates a
(http://www.cibiv.at/~greg/haploviewer). Individuals were assigned species-tree (Yang and Rannala, 2014), using the species tree estimated
to populations following STRUCTURE/Geneland results (see below). in StarBEAST2 as the guide tree. Each run consisted of a burn-in phase of
10,000 iterations followed by a sampling phase of 400,000 iterations,
2.3. Bayesian genetic assignment with a frequency of 4 iterations per sampling. Posterior probabilities ≥
0.95 were considered highly supported.
Because missing data for POMC was considerably high (33% out of
144 individuals), we only used cmyc2, Fib7 and Tyr nuclear genes for

3
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

2.4. Genetic diversity and historical demography 2.5. Species tree and divergence times

To assess genetic diversity, we estimated haplotype diversity (Hd) We estimated a dated species tree using all loci, mitochondrial and
and per-site nucleotide diversity (Pi) using DnaSP (Librado and Rozas, nuclear, for all samples in StarBEAST2 v0.15.13 (Ogilvie et al., 2017) as
2009). We also applied the neutrality tests of Tajima’s D (Tajima, 1989), implemented in BEAST v2.6.6 software. We grouped individuals into
Fu’s Fs (Fu, 1997) and Ramos-Onsins and Rozas’s R2 (Ramos-Onsins and eight populations that were identified by STRUCTURE/Geneland and
Rozas, 2002) to detect significant deviations from the null hypothesis of supported as distinct evolving lineages with significant posterior prob­
neutral evolution and constant population size. We estimated signifi­ abilities (≥0.95) in BPP. Because StarBEAST2 cannot model migration
cance levels of neutrality tests through 10,000 coalescent simulation or introgression, we also removed samples that showed incongruences
replicates in DnaSP. Because neutrality tests were significant for some between nuclear assignment relative to its mitochondrial cluster (eight
populations (Table 1), we then performed the multilocus coalescent- individual’s total; see Fig. 1). To run the species-tree we also set mean
based extended Bayesian skyline plot (EBSP; Heled and Drummond, substitution rates of 16S and COI as specified in the previous section.
2008) implemented in Beast v2.6.6 (Bouckaert et al., 2019) to estimate Substitution rates of the nuclear genes were coestimated by the program.
changes in effective population sizes over time. We used all loci, mito­ We assumed a relaxed clock model with uncorrelated lognormal distri­
chondrial and nuclear (except for POMC), and the 16S substitution rate bution for the mitochondrial genes, strict clock for the nuclear genes,
as a reference (mean = 5.42 × 10-3 substitutions/million year; SD = and a birth–death model of diversification as tree prior (see Table S3 for
1.03 × 10-3) following recent estimates for Leptodactylus (Fouquet et al., details about substitution models). Because we have no prior informa­
2012). The COI mitochondrial segment proportionally accumulates tion on ancestral effective population sizes, we used the Linear with
nucleotide mutation at a rate twice as fast as the 16S gene (Vences et al., Constant Root as populational model prior. We ran the MCMC chain for
2005; Lyra et al., 2017). Therefore, we also set the COI mean substitu­ 200,000,000 iterations, sampling every 20,000 iterations, and discarded
tion rate as 1.0 × 10-2, and relaxed the clock so that the HDP 95% the initial 25% iterations as burn-in. To check for convergence among
confidence interval prior ranged from 6.0 × 10-3 to 2.0 × 10-2 sub­ runs, we performed three independent analyses using different seed
stitutions/site/Mya (M = 0.01; S = 0.25). For both clock calibrations, we numbers and combined logs with LogCombiner (BEAST package). We
employed a strict molecular clock and uniform priors to set the HPD checked for run convergence by visually checking ESS (>200) values in
interval. We implemented the same substitution models provided in the log files using Tracer v1.7. We summarized the species tree in Tree
Table S3. We ran 50,000,000 MCMC simulations sampled every 5,000 Annotator, which was drawn in FigTree v1.6 (Rambaut, 2014). To
chains. We checked for run convergence using Tracer 1.7 (Rambaut further check for concordance of all independently sampled gene trees,
et al., 2018) by visually inspecting effective sample size (ESS > 200). we visualized the species tree with DensiTree (Bouckaert, 2010).

Table 1
Genetic summary statistics for each locus for six lineages of the Leptodactylus latrans complex (lucSC and payN not included because of low sampling). N: sample size;
bp: fragment length in base pairs; S: number of polymorphic sites; H: number of haplotypes; Hd, haplotype diversity; pi: nucleotide diversity. *Significant p value for
neutrality tests (boldface).
Locus Population Summary statistics Neutrality tests

N bp S H Hd pi D R2 Fs

16S mtDNA total 125 348


payS 15 – 1 2 0.133 0.000 1.159 0.249 − 0.649
lucN 40 – 7 6 0.432 0.004 − 0.314 0.103 − 0.162
lucS 25 – 16 11 0.827 0.008 − 1.325 0.075* ¡3.509*
latN 12 – 7 7 0.894 0.005 − 0.946 0.103* ¡2.976*
latS 16 – 3 4 0.442 0.002 − 1.002 0.112* ¡1.415*
par 17 – 6 7 0.596 0.002 ¡1.825* 0.078* ¡4.799*
COI mtDNA total 140 507
payS 19 – 16 6 0.725 0.014 2.093* 0.22 4.086
lucN 42 – 39 22 0.916 0.014 − 0.838 0.0852 ¡5.352*
lucS 27 – 42 17 0.892 0.016 − 1.001 0.087 − 3.196
latN 17 – 20 13 0.949 0.012 − 0.243 0.129 ¡4.104*
latS 18 – 30 13 0.961 0.010 − 1.687 0.092 ¡4.36*
par 17 – 6 6 0.654 0.002 − 1.631 0.098* ¡2.848*
Fib7 (nuDNA) total-phased 210 359
payS 26 – 3 4 0.702 0.004 1.791 0.227 1.049
lucN 62 – 6 6 0.594 0.002 − 0.809 0.072 − 1.296
lucS 32 – 5 5 0.542 0.004 0.119 0.129 0.135
latN 32 – 3 3 0.365 0.001 − 0.829 0.13 0.014
latS 32 – 6 6 0.806 0.005 0.399 0.139 0.004
par 26 – 4 5 0.625 0.002 − 0.782 0.096 − 1.576
Tyr (nuDNA) total-phased 246 510
payS 30 – 4 4 0.490 0.001 − 0.855 0.0842 − 0.532
lucN 74 – 17 24 0.948 0.008 0.524 0.121 ¡8.399*
lucS 46 – 11 12 0.831 0.006 0.884 0.147 − 1.517
latN 32 – 2 3 0.433 0.001 − 0.190 0.13 − 0.104
latS 32 – 5 5 0.631 0.002 − 0.805 0.088 − 0.93
par 32 – 5 6 0.782 0.003 0.123 0.129 − 0.79
cmyc2 (nuDNA) total-phased 276 852
payS 36 – 4 4 0.410 0.001 − 1.123 0.101 − 0.968
lucN 84 – 9 12 0.501 0.001 − 1.764 0.039* ¡10.123*
lucS 52 – 5 9 0.729 0.002 0.424 0.13 − 2.839
latN 34 – 5 6 0.670 0.001 − 0.567 0.108 − 1.606
latS 36 – 6 8 0.687 0.001 − 0.952 0.092 ¡3.109*
par 34 – 3 4 0.604 0.001 0.339 0.14 0.102

4
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

Fig. 1. Inter and intraspecific genetic di­


versity among members of the Leptodactylus
latrans species complex. (a) Mitochondrial
gene tree recovered in BEAST and genetic
breaks found by STRUCTURE (STR) and
Geneland (GNL) for each species using nu­
clear data only. The first column refers to
species delimited in the taxonomic review of
Magalhães et al. (2020). Values on nodes
indicate posterior probabilities, while black
circles denote posterior probability = 1.0.
Scale indicates rate of base substitutions per
site. Asterisks indicate lineages validated by
BPP with probabilities above 0.95 combining
mitochondrial and nuclear loci. Lineages are
named as north (N) and south (S) pop­
ulations relative to the geographic break,
while lucSC is named after the Brazilian state
it occurs (Santa Catarina). (b) Haplotype
networks for the four nuclear genes, cmyc2,
Tyr, Fib7, and POMC (clusters and haplotype
colors match those clusters validated by
BPP). Numbers within circles refers to
haplotype frequencies. Branches without
dots represent a single mutational step. Dots
along haplotype networks indicate addi­
tional mutational steps for branches with
more than one mutation.

2.6. Ecological niche models (delta AICc = 0). Variable importances, AUC values, and FC and RM
combinations of models are resumed in Table S5 of Supplementary
To characterize the environmental niche of all deeply divergent Material.
species within the L. latrans complex (delimited by Magalhães et al., Finally, we projected present-day ENMs into bioclimatic variables
2020), we assembled a minimum of 30 localities for each species predicted for two different past scenarios, Last Interglacial (LIG; ca. 130
(Proosdij et al., 2016). We used as occurrence point data all localities kya) (Otto-Bliesner et al., 2006) and Last Glacial Maximum (ca. 21 kya)
associated with molecular-vouchered individuals plus the database of from the Model for Interdisciplinary Research on Climate (MIROC;
species occurrences based on museum records assembled by Magalhães Hasumi and Emori, 2004). Because Pleistocene climatic oscillations
et al. (2020). We obtained a total of 80 occurrence points for L. latrans, directly shaped landscape and biomes ranges (Carnaval et al., 2014;
192 for L. luctator, 30 for L. payaya, and 30 for L. paranaru. Gehara et al., 2017b), we expect to correlate past environmental niche
We generated environmental niche models (ENM) using bioclimatic scenarios to populational demographic changes among species/lineages
data from WorldClim (Hijmans et al., 2005; http://www.worldclim.org/ in the L. latrans species complex.
) at a spatial resolution of 2.5 arc-minutes. The 19 “Bioclim” layers
(Table S4) reflect aspects of precipitation and temperature. Prior to 2.7. Niche overlap/divergence
analyses, raster layers were clipped to enclose a reasonable extent (-70
to − 34 degrees of longitude and − 41 to − 5 degrees of latitude) To evaluate the degree of niche overlap among species in the Lep­
considering L. latrans complex broad distribution along eastern South todactylus latrans complex and to test whether diversification can be
America (see Magalhães et al., 2020). Then, to avoid model over- associated to niche divergence, we used a Principal Component Analysis
parameterizations (Rissler et al., 2007), we excluded highly correlated of the environmental space (PCA-env) approach (Broennimann et al.,
bioclim variables using variance inflation factor (VIF; threshold > 10) 2012) using ecospat R package (Di Cola et al., 2017). For such purpose,
with the usdm R package (Naimi et al., 2014). After this procedure, nine we calculated the niche overlap for each pair of species with Schoener’s
bioclim predictors were retained: mean temp. diurnal range (Bio2), D metric (Schoener, 1968), and assessed if the observed D is more similar
isothermality (Bio3), mean temp. of wettest quarter (Bio8), mean temp. than random expectation with pairwise niche similarity tests (Warren
of driest quarter (Bio9), precip. of wettest month (Bio13), precip. of et al., 2008; Broennimann et al., 2012). We tested the niche conserva­
driest month (Bio14), precip. seasonality (Bio15), precip. of warmest tism hypothesis using 1,000 permutations to assess significance (P <
quarter (Bio18), and precip. of coldest quarter (Bio19). We generated 0.05), and allowing random shifts simultaneously of both niches (rand.
ENMs with Maxent 3.3 algorithm (Phillips et al., 2006) and used the type = 1 option). We tested the niche conservatism hypothesis using the
ENMeval package in R (Muscarella et al., 2014) for tuning Maxent “greater” option. If the observed D-values are significantly different from
models. To rank and select models, we used default parameters of random expectation, the hypothesis of niche similarity is accepted. The
feature classes (fc = “L”, “LQ”, “H”, “LQH”, “LQHP”, “LQHPT”) and available environmental space was resumed by the first two PCs, and
regularization multipliers (RMvalues = seq(0.5, 4, 0.5)) in ENMEval. because this group is composed of large-sized species with an expected
The model performance was based on the area under the receiver- high dispersion rate, we used the minimum convex polygon surrounding
operating characteristic curve (AUC) and Akaike information criterion the original occurrence records within a buffer of 300 km of each species

5
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

to avoid drawing background values from areas without species occur­ lineages are not necessarily allopatric, there is a strong pattern of
rence but also as a likely environmental space available. D-values range geographic structure at populational scale as well. Interestingly,
from 0 to 1, with 0 representing no niche overlap and 1 corresponding to L. paranaru (par) seems to share haplotypes at a higher proportion with
identical niches between the two compared groups. Also, the degree of the southern population of L. latrans (latS), than southern and northern
niche overlap can be categorized as no or very limited overlap (D populations of L. latrans are sharing between themselves (Fig. 1).
varying from 0 to 0.2), moderate (0.2–0.4), high (0.6–0.8), and very
high niche overlap (0.8–1.0) (Rödder and Engler, 2011). 3.2. Bayesian genetic assignment
In order to examine niche divergence at a finer scale, we also
assembled datasets at populational level for species that showed genetic The population assignment test of STRUCTURE identified an optimal
and geographic population structure (identified in our phylogeographic value of two clusters for all the four species in the Leptodactylus latrans
analyses) and have a minimum of 15 localities sampled (only L. latrans complex (Fig. 1), totaling eight populations. Conversely, Geneland
and L. luctator fit these above-mentioned requirements). For the niche identified a total of nine genetic clusters with posterior probability over
characterization at the populational scale, we only considered as 0.65 (Supplementary Fig. 1). The results of both assignment tests were
occurrence data points molecular voucher specimens (depicted in overall congruent, except that Geneland showed one additional break
Fig. 3a): we compiled 42 localities for the L. luctator northern lineage within L. payaya (splitting populations distributed southwards of the São
(lucN) and 28 for L. luctator southern lineage (lucS), and 19 localities for Francisco River in two; Supplementary Figs. 1 and 2) and L. luctator
the L. latrans northern lineage (latN) and 18 for L. latrans southern (splitting apart the Serra Geral population in the Brazilian State of Santa
lineage (latS). We applied the same methodological procedure for spe­ Catarina; Figs. 1 and 2), that were not validated by STRUCTURE.
cies and population scales. We applied the same background of the Moreover, Geneland did not detect a genetic break within L. paranaru
species for its respective populations. ENMs and niche overlap analyses lineage, as identified by STRUCTURE (Fig. 1).
were performed in R v4.0.2 (www.r-project.org). All of the four prior combinations and species delimitation algorithm
(0 or 1) used in BPP consistently supported with high pp. (>0.95) the
3. Results existence of six lineages (latN, latS, lucN, lucS, lucSC, and par) within
the Leptodactylus latrans complex, but instead showed incongruences
The final alignment for the 16S and COI mitochondrial genes com­ regarding the delimitation of northern and southern lineages of
prises 350 bp (37 haplotypes) and 507 bp (77 haplotypes), respectively. L. payaya (payN and payS; Table S6 in Supplementary material). For
Alignments of the nuclear gene segments included 852 bp of cmyc2 (43 instance, only the runs with theta prior assuming large population sizes
haplotypes), 359 bp of FIB7 (29 haplotypes), and 510 bp of Tyr (54 [IG(3,0.2)] lumped the lineages payN and payS into a single cluster with
haplotypes), and 407 bp of POMC (7 haplotypes) (Table 1; Fig. 1b). significant probability > 0.95, while theta priors assuming small pop­
Among 144 sequenced individuals, the percentage of missing data for ulation sizes [IG(3,0.002)] support these two lineages with significant
Fib7, Tyr, cmyc2 and POMC nuclear genes were 12%, 6%, 1% and 33%, probability (PP > 0.98). The additional break within payS validated by
respectively, and 11% for 16S mitochondrial gene (all individuals had Geneland was not supported in any model. Moreover, all prior combi­
COI sequenced). Moreover, both mitochondrial gene tree topology and nations lumped the two lineages within L. paranaru (identified only by
monophyly of the Leptodactylus latrans complex (Fig. 1) agree with STRUCTURE) into a single species with significant support (pp > 0.95),
previous proposals for the L. latrans species group (Magalhães et al., except when nuclear loci are analyzed separately under small theta
2020). priors (Table S6). However, sample size and number of loci used may
impact BPP inferences by splitting populations from the same species
3.1. Haplotype network genealogies and genetic diversity (Zhang et al., 2014), which might be the case. Indeed, this break within
L. paranaru did not exhibit cohesive phylogenetic nor geographic
We found high nucleotide and haplotype diversity within COI, while structure (see Supplementary Fig. 3) and likely represents a STRUC­
summary statistics revealed slightly lower diversity within 16S and TURE artifact, given the impossibility of validating K = 1 by means of
nuclear genes when compared to COI (Table 1). Nevertheless, haplotype Evanno’s methods (Janes et al., 2017). Therefore, BPP analyses vali­
networks showed a relatively high genetic variation in nuclear gene dated a total of seven or eight lineages with high posterior probability
haplotypes within the Leptodactylus latrans species complex (Fig. 1b), depending on theta prior we employ.
except for POMC for which only seven haplotypes were recovered. Despite mitochondrial and nuclear genes overall congruence, we
Moreover, among FIB7, POMC and Tyr nuclear genes only a single detected some incongruences between population assignment based on
haplotype is shared between at least two of the four species delimited in nuclear loci and mitochondrial clade within lucN-lucS and latN-latS
Magalhães et al. (2020) taxonomic review, and haplotype sharing individuals, likely as a result of gene flow or introgression. For
occurred in a slightly higher proportion at the populational level. For instance, six individuals clustered within lucS mitochondrial lineage
instance, we observed 43 haplotypes for cmyc2, of which four are shared were assigned to lucN by the nuclear genes-based assignment tests,
between at least two distinct lineages, 54 haplotypes for Tyr (nine while one individual from lucN was assigned to lucS lineage. A single
shared), 29 haplotypes for FIB7 (eight shared), and seven haplotypes for individual from latN mitochondrial lineage was assigned to the latS by
POMC (four shared). Such amounts of incomplete lineage sorting could nuclear genes. Their final populational assignment was based on the
indicate that lineages diverged under the presence of gene flow, lineages nuclear loci. Interestingly, the southern population of L. payaya was not
diversified in a short timespan relative to population sizes or are a result recovered as reciprocally monophyletic in the mitochondrial genes as
of secondary contact (Fig. 1). Additionally, the geographic distribution some individuals are more related to the northern population clade (pp.
of species in the L. latrans complex is mostly non-overlapping with very = 0.8; Fig. 1). Nevertheless, these mitochondrial clades are genetically
few sympatry/syntopy zones occurring along their distribution (Mag­ structured, suggesting that a past event of mitochondrial introgression
alhães et al., 2020). Likewise, we observed a similar non-overlapping might have occurred.
spatial pattern within lineages, except for one locality in the Brazilian
state of Minas Gerais where northern and southern lineages of L. latrans 3.3. Historical demography
(latN and latS, respectively) and the northern lineage of L. luctator (lucN)
were sampled in sympatry, and another locality at Chapada Diamantina The neutrality tests values were not consistent among algorithms,
in central Bahia State where lucN and southern L. payaya lineage (payS) genes and within lineages. For instance, neutrality test values were
lineages are sympatric. In both cases, we found no evidence of admixture significant for mitochondrial genes (except for payS lineage), which
among individuals in any gene (mitochondrial and nuclear). Although instead were mostly not significant for nuclear genes (Table 1). The

6
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

Fig. 2. Extended Bayesian skyline plot


(EBSP) showing changes in effective popu­
lation size through time for lineages in the
Leptodactylus latrans species complex based
on mitochondrial and nuclear genes. Y-axis
shows the effective population size with
respective mean (dashed lines) and 95%
highest posterior density limits (the sur­
rounding gray area delimited by the contin­
uous lines). The x-axis shows time in millions
of years as time goes backwards from left to
right. The red and blue vertical bars along all
plots indicate the Last Interglacial (LIG, ca.
130 kya) and Last Glacial Maximum (LGM,
ca. 21 kya) periods, respectively. (For inter­
pretation of the references to colour in this
figure legend, the reader is referred to the
web version of this article.)

inconsistency among nuclear and mitochondrial genes may be attrib­ (a lineage restricted to the Chapada Diamantina ecoregion).
uted to differences in mutation-accumulation rates, which are overall
more accelerated in the mitochondrial genome (Lynch, 2007). There­ 3.4. Species tree and divergence time
fore, abrupt changes in effective populational sizes are more efficiently
detected by mitochondrial genes, especially when these changes Both StarBEAST2 and BPP recovered the same species tree topology
occurred in a recent timespan, which might be the case. Nonetheless, with high probabilities (PP > 0.98) for all internal nodes (see Fig. 3b).
these results support signs of recent demographic expansion. According Although the posterior probability of the best species tree model varies
to Ramos-Onsins and Rozas (2002), Fs and R2 tests are more robust in according to genes employed and prior combinations in BPP (Table S6),
detecting events of demographic expansion, and Fs is more suitable for they all agree with StarBEAST2 species tree topology. As expected,
larger samples while R2 is more suitable for smaller samples. significant species tree models were only retrieved when mitochondrial
The EBSP analysis revealed a pronounced increase in effective pop­ and nuclear loci are analyzed together. The DensiTree representation of
ulation size for all populations distributed within the AF boundaries the species tree also shows that most gene trees sampled agree with the
such as latN, latS, lucN and Par at around 130,000 years ago. Moreover, maximum clade credibility tree (see Supplementary Fig. 4), confirming
a very subtle increase in effective population size was detected for lucS that the recovered topology is highly congruent among inferences. The
lineage after LGM (Fig. 2). Conversely, EBSP revealed no changes on resulting tree from StarBEAST2 shows that the origin of the Leptodactylus
effective population sizes during the last 300,000 years for payS lineage latrans species complex and the time to L. luctator, L. latrans, and L.

7
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

Fig. 3. (a) Geographic distribution along


eastern South America of all genetic samples
from all eight supported lineages belonging
to the Leptodactylus latrans species complex;
(b) Time-calibrated species tree of lineages in
the L. latrans species complex estimated with
StarBEAST2. Terminals represent the genetic
breaks found by Geneland and STRUCTURE
and validated by BPP. Values above nodes
indicate Bayesian posterior probabilities,
while values below indicate mean divergence
time estimated in StarBEAST2. Asterisks
indicate posterior probabilities values = 1.0.
Scale indicates time in millions of years
(Mya). The main rivers that could be related
to populational genetic breaks are also
depicted. Country acronyms: ARG:
Argentina, BRA: Brazil, PAR: Paraguay, URU:
Uruguay.

paranaru most recent common ancestor are both within the boundaries within L. latrans, L. luctator (except for lucSC) and L. payaya showed a
of late Miocene around 5.5 Mya (95% HPD = 2.8–8.5 Mya) and 3.3 Mya synchronous diversification during late Pleistocene around 200–300 kya
(95% HPD = 1.79–5.1 Mya), respectively. Diversification between the (Fig. 3b). Because nuclear and mitochondrial genes are mostly recipro­
AF species (L. latrans and L. paranaru) falls within the transition between cally monophyletic (as showed by haplotype networks) and StarBEAST2
Plio-Pleistocene boundaries around 1.7 Mya (95% HPD = 0.6–2.9 Mya). topology was recovered with high node supports, lineages likely diver­
Conversely, all other splitting events within species started during the sified under a scenario where gene flow is absent or very restricted
middle-to-late Pleistocene about 600 kya (HPD 95% = 0.3–1.0 Mya) (Degnan and Rosenberg, 2009).
with the early divergence within L. luctator (Fig. 3b). Moreover, lineages

Fig. 4. Ecological niche models constructed in Maxent for all four species in the Leptodactylus latrans complex projected to the present, last glacial maximum (LGM),
and last interglacial (LIG) periods. Hot colors indicate high habitat suitability, while cold colors indicate low habitat suitability.

8
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

3.5. Ecological niche models and niche comparisons evidenced by niche suitability displacement when ENMs are projected to
LIG (ca. 130 Kya; Fig. 4).
The predicted suitable areas for species in the Leptodactylus latrans Moreover, both latN and latS populations showed signals of popu­
complex based on the ENMs matched their actual occurrences closely lation expansions about 130,000 years ago, which corroborates the
(Fig. 4). The mean value of the area under the operating receiver curve hypothesis that AF associated taxa expanded rather than retracted
(AUC) was similar between the four species (>0.90 for all species), during LGM (Leite et al., 2016). This is especially interesting given that
indicating better than random predictions (0.5 = random, 1 = all populations distributed along the AF coast (L. paranaru included)
maximum). Considering pairwise comparisons between present pro­ showed signs of synchronous population expansion around 130,000
jections of species ENMs, models consistently did not predict over­ years ago (Fig. 2). Additionally, lineages distributed south of the
lapping geographic areas well (e.g., the environmental niche of one climatically stable areas (e.g., lucN and latS) proposed by Carnaval and
species projected to the other). We recovered a similar pattern of niche Moritz (2008) exhibited similar levels of genetic diversity in comparison
suitability influenced by Pleistocene climatic oscillations among all four to northern AF lineage (latN). This indicates that the LGM period did not
species in the L. latrans complex (Fig. 4). For instance, we observed a have significative effects on population size changes nor environmental
trend of habitat displacement toward northern areas during the LGM stable areas predicted genetic diversity as expected by the CM model
and towards southern areas during the LIG period. Moreover, we found (Carnaval and Moritz, 2008), a pattern also observed for another frog
more fragmentation of suitable habitats during the LIG than in the LGM, complex (Thomé et al., 2014) and birds (Cabanne et al., 2016).
especially for AF endemic species (e.g., L. latrans and L. paranaru). Among all species in the Leptodactylus latrans complex, only L. par­
Interestingly, we observed a significant expansion of niche range for AF anaru did not exhibit signs of populational genetic structure, but this
species towards the exposed continental shelves during the LGM in species is restricted to a narrow geographic area along low altitudinal
comparison to LIG. sites (bearing sea-level) located east to the Serra do Mar Mountain range
The niche overlap between pairwise comparisons among all four (Fig. 3). Our divergence time estimates indicate that the AF endemic
species in the L. latrans complex ranged from none to very limited, with species (L. paranaru and L. latrans) diverged during the Plio-Pleistocene
D-values ranging from 0 to 0.18 at the interspecifc level (Fig. 5, Table 2), transition, around 1.7 Mya. The phylogenetic break between these two
suggesting a general pattern of niche divergence. For the six pairwise species occurs within Santos Basin in São Paulo State, a region where lies
combinations, only L. luctator/L. latrans showed signals of niche the Continental Rift of Southeastern Brazil (Vieira et al., 2015). Phases
conservatism, but with low value of niche overlap (P < 0.05, D = 0.18) of tectonic reactivations of ancient structural faults have been promot­
(Fig. 5). Thus, most pairwise combinations suggest that the niches of ing changes along this continental rift (Souza, 2015), and allied with
species were not more similar than random expectations, rejecting the sea-level changes transgressions/regressions related to Quaternary
hypothesis of niche conservatism. This pattern was also observed at interglacial periods (Siddal et al., 2010; Hansen et al., 2013), may have
populational scale as similarity tests indicated niche divergence between isolated a once continually coastal distributed lineage. The diversifica­
lucN and lucS lineages and for latN and latS lineages (Fig. 5g,h), with D- tion of a highly endemic fish fauna is mostly associated with such neo­
values of 0.07 and 0.20, respectively (P > 0.05; Table 2). tectonic activities (Ribeiro et al., 2006), with one of them (the Bertioga
Fault; see Souza, 2015) matching our mean divergence time estimates
4. Discussion for the split between L. paranaru and L. latrans. Additionally, the current
geographic distribution of L. paranaru lies within an area where neo­
4.1. Timing of diversification and biogeographic patterns tectonic barriers (e.g., the Guapiara lineament and the Cubatão shear
zone; Saadi et al., 2002) coincide with phylogeographic breaks among
We identified that three of the four species delimited in Magalhães several AF-associated taxa (e.g., Grazziotin et al., 2006; Thomé et al.,
et al. (2020) exhibit considerable genetic diversity with geographically 2010; Menezes et al., 2017). However, the large confidence intervals of
structured populations that diverged during the late Pleistocene, divergence time estimates (HPD 95% = 0.6 to 2.9 Mya) and the uncer­
totaling eight independently evolving lineages. Considering popula­ tainty in establishing a single tectonic barrier to gene flow, hampered us
tional phylogeographic breaks, all delimitation analyses validated two to clearly distinguish whether an ecological or vicariance based process
genetically structured populations belonging to Leptodactylus latrans explains this phylogenetic break. Although niche comparisons indicate
(Fig. 1). The break between northern (latN) and southern (latS) pop­ that environmental niches of L. paranaru and L. latrans are not similar
ulations matches the Doce River break, as predicted by CM model for AF (Table 2; Fig. 5), patterns of ecological isolation can also be a by-product
associated taxa (Carnaval and Moritz, 2008). The Doce River has been of allopatric isolation mostly because L. paranaru lineage has been
considered a barrier to gene flow in previous phylogeographic studies affected by at least three neotectonic pulses. Interestingly, latS and L.
(Costa and Leite, 2012). However, its relative importance as a major paranaru lineages still share haplotypes at a high proportion. Even
barrier to gene flow is questionable because it runs through an area of though our results did not support sympatry between these lineages
noticeable environmental/climatic shift (Carnaval and Moritz, 2008; through genetic sampling, a secondary contact would be possible due to
Carnaval et al., 2014). Hence, it is likely that vicariance is driven by range expansion in overlapping regions during LGM (Fig. 4), as L. latrans
climatic isolation, rather than river-associated (Carnaval et al., 2014), (latS lineage) was sampled about 25 km northwards from the distribu­
although it has been suggested that the river could have a secondary role tion limit of L. paranaru along São Paulo State coastal zone (Magalhães
(limiting dispersal; Colombi et al., 2010; D’Horta et al., 2011). Because et al., 2020). As proposed by the ENMs (Fig. 4), this lineage remained
species in the L. latrans complex are all large and likely not constrained restricted to a narrow geographic patch between northern Paraná and
to disperse over long distances, specially across river margins (Fonte São Paulo States, a region of climatic refugia with high phylogeographic
et al., 2019), a climatic-driven isolation scenario seems more reasonable endemism index (Carnaval et al., 2014), showing recent signs of popu­
than river-associated one, especially because sedimentary evolution of lational expansion as sea-levels continuously regressed during the last
Doce River (Cenozoic) vastly precedes the diversification between these 130,000 years (Siddal et al., 2010; Hansen et al., 2013).
two populations (Mello et al., 1999). Moreover, if rivers were strong Leptodactylus luctator exhibited the highest uncovered diversity
barriers to gene flow, we would expect that other main rivers along among species in the L. latrans complex and three genetically structured
Bahia State (e.g., Jequitinhonha River) and Rio de Janeiro States (e.g., populations are recognized. Among these, individuals from a high-
Paraíba do Sul River) would also have promoted phylogeographic elevation site at Serra Geral region in the Brazilian State of Santa Cat­
breaks within L. latrans, what is not the case here. Therefore, it is likely arina (lucSC) diverged from the sister lineages (lucN + lucS) during the
that habitat fragmentation caused by Pleistocene climatic oscillations is middle to late Pleistocene around 600,000 years ago. Recently, both
the primary mode for lineage diversification within this species, also Carnaval et al. (2014) and Barros et al. (2015) predicted that a narrow

9
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

Fig. 5. Niche comparisons among species/populations in the Leptodactylus latrans complex in the environmental space. The twofirst principal components axes of the
PCA-env under present climatic scenario represent niche characteristics of each lineage pair. Graphs represent the environmental space occupied by both lineages,
with darker cells showing the highest density of occurrences and available background in lines. Interspecific comparisons between (a) L. latrans vs. L. luctator, (b) vs.
L. paranaru, and (c) vs. L. payaya; (d) L. luctator vs. L. paranaru, (e) vs. L. payaya; (f) L. paranaru vs. L. payaya. Intraspecific comparisons between northern and
southern lineages of (g) L. latrans and (h) L. luctator are also shown. The contributions of the climatic variables to the two axes of the PC analysis and the percentage of
the variation explained by the two axes are also provided. See methods for environmental variables.

10
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

Table 2 sharing with species in the L. latrans complex (being all exclusive to this
Niche comparisons at inter and intraspecific levels for Leptodactylus latrans species), and a unique and distinct advertisement call among species in
complex. Pairwise niche overlaps (Schoener’s D metric) and P-values of niche the L. latrans complex (Magalhães et al., 2020). This species showed a
similarity tests considering the hypothesis of niche conservatism. In boldface populational break that split north (payN) and south (payS) populations
intraspecific comparisons. *Significant P-values (boldface). around where the São Francisco River currently flows. The São Francisco
Comparison D Similarity (P-value) River lower course previously flowed in a different direction northward
latrans vs luctator 0.183 0.005* towards the equatorial Atlantic Ocean and drastically shifted towards
latrans vs paranaru 0.11 0.092 the southeast and to the eastern Atlantic Ocean coast about approxi­
latrans vs payaya 0.034 0.465 mately 450,000 years ago (Mabesoone, 1994; Nascimento et al., 2013).
luctator vs paranaru 0.076 0.127
The shift in the São Francisco River course precedes the divergence time
luctator vs payaya 0.029 0.544
paranaru vs payaya 0 1 confidence interval estimated for the split between payN and payS
latN vs latS 0.205 0.235 populations (mean = 0.2 Mya; HPD 95% = 0.056–0.39 Mya), indicating
lucN vs lucS 0.072 0.314 that a riverine barrier hypothesis does not explain the diversification of
L. payaya lineages (as previously reported for other taxa; Nascimento
et al., 2013; Bruschi et al., 2019; Lanna et al., 2020) and might rather
patch of higher altitudinal sites across Paraná and Santa Catarina States
have had a secondary role in limiting dispersal (as proposed for another
acted as forested microrefugia and played an important role in main­
frog species; Thomé et al., 2021). On the other hand, Pleistocene
taining genetic diversity locally. Although not a forest-dependent taxon,
interglacial cycles continuously promoted aridification of lowland
our results corroborate the existence of a deeply divergent lineage
Caatinga areas (Auler and Smart, 2001) with a prominent expansion of
within L. luctator that was only sampled from Serra Geral region in Santa
Caatinga and associated dry-adapted taxa during the last 260,000 years
Catarina State. Accordingly, Peçanha et al. (2017) identified that cryptic
(Gehara et al., 2017b). However, several recurrent wetter periods
rodent lineages associated to open areas diversified around this same
occurred within the Caatinga Biome during the middle to late Pleisto­
region, with divergence time estimates falling within the middle to late
cene (Wang et al., 2004; Auler et al., 2004) promoting the expansion of
Pleistocene boundaries.
forest corridors that allowed species to disperse, especially along
The ENMs consistently identified two main core areas that match the
northern and central Bahia State. These wetter pulses date back to
geographic ranges of lucS (around northeastern Argentina/southern
middle Pleistocene about 900,000 years ago (Auler et al., 2004), which
Paraguay) and lucN (around São Paulo and Minas Gerais State) lineages.
intensified during the last 210,000 years (Wang et al., 2004). This is
These core areas are relatively fragmented during both LGM and LIG
interesting because the current location of the lower São Francisco River
periods as ENMs predicts low environmental suitability around the
course was pointed in the ENMs as an area of high suitability for this
higher elevation sites within the AF region comprising the interior of São
species during both LIG and LGM, which is substantiated by evidences of
Paulo, Paraná and Santa Catarina States (Fig. 4). For instance, previous
wetter pulses within the Caatinga during these periods (Auler and Smart,
palynological studies showed that highlands in southern Brazil exhibited
2001). Considering that L. payaya is not a dry-adapted species, it is likely
a relatively dry climate with a longer annual dry season and a cold
that these early wetter pulses favored populational dispersion outside its
climate with frosts during glacial cycles (Behling and Lichte, 1997).
core distribution in central Bahia followed by isolation when Caatinga
Only during late Holocene, the grassland fields were replaced by forests
aridification intensified in late Pleistocene (Auler and Smart, 2001),
(related to increasingly moister climate change; Behling, 1995), which
which instead favored populational expansion of dry-adapted taxa
expanded into the highlands of southern Brazil (Behling, 2002), likely
(Gehara et al., 2017b). Therefore, both higher elevation areas of Per­
favoring lineages to disperse and to establish secondary contact. Indeed,
nambuco and Chapada Diamantina highlands (which are embedded
we only observed mitochondrial introgression and sympatry between
within a semi-arid landscape; Vieira et al., 2015) could have acted as
lucN and lucS lineages along AF and Chaco/Pampas biomes transitional
climatic refugium during dryness periods, and diversification was driven
areas located around southern Brazil and northeastern Argentina
by climatic oscillations. This further explains why population size of
(Fig. 3). This ecoregion is also referred to as Southern Cone Meso­
L. payaya southern lineage remained constant (as show by EBSP anal­
potamian Savanna (Olson et al., 2001), and is well known for repre­
ysis) or experience bottle necks as indicated by significant Tajima’s D
senting a transitional area between biotas of the Chaco and the Atlantic
positive values for the COI mitochondrial gene (Table 1).
Forest (Giraudo and Arzamendia, 2017). In addition, in this region, the
Paraná River has suffered historical fluctuations, particularly during the
4.2. Patterns of ecological isolation
Quaternary (Popolizio, 2004), that may have acted as a geographical
barrier limiting gene flow between southern and northern L. luctator
Divergence in ecological niches seems to be an important factor in
lineages, as suggested for mammals and birds (e.g., Márquez et al., 2006;
shaping the diversification processes of South American butter frogs.
Kopuchian et al., 2020). Nevertheless, because there were no major
Most comparisons for both inter and intraspecific levels showed that
geographic barriers to gene flow at the time lineages divergence (from
lineages tended to occupy different environmental conditions and niche
180,000 to 560,000 years ago) and niche comparisons support envi­
evolution presented none or low niche overlap. As general rule, phylo­
ronmental divergence (Table 2), a climate/ecological based process
genetic sister species often exhibit differences on traits that are apparent
likely explains the patterns of genetic diversification within L. luctator.
responses to differing ecological conditions, revealing a primary mode of
Moreover, EBSP analyses indicated a subtle, but noticeable (as sup­
ecological isolation on driving lineage divergence and speciation (Mayr,
ported by neutrality testes; Table 1) sign of populational expansion
1947; Shafer et al., 2013). Nevertheless, except for some unique intra­
around the LGM for both lucN and lucS lineages. Conversely to other AF
specific chromatic variations, external morphology among species in the
species (e.g., L. latrans [latN, latS] and L. paranaru), EBSP analysis
Leptodactylus latrans complex is overall homogeneous (Magalhães et al.,
supports a post-LGM population expansion for this cold-associated taxon
2020). Moreover, an extensive mitochondrial DNA barcoding sampling
(lucN and lucS) despite ENMs predictions of range expansion towards
(with more than 300 sequenced individuals) detected relatively few
continental shelves or Cerrado since LIG, while some physical barriers
sympatry zones among species in the L. latrans complex, despite their
(e.g., Serra do Mar Mountain range) or physiological limitations could
broad geographic occurrence area across South America. This is espe­
have constrained population viability during Pleistocene glacial cycles.
cially interesting considering that ENMs among species pertaining to the
Leptodactylus payaya is the first species to split within this complex of
L. latrans complex are mostly not overlapping under current climatic
frog species and originated at about 5.5 Mya during the late Miocene.
conditions (Fig. 4). Thus, these sister species may be adapted to different
This old divergence time reflects in a complete absence of haplotype
environmental gradients or have allopatric/non-overlapping

11
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

distributions enforced by their similar behaviors (e.g., calls, microhab­ Writing – review & editing. Mariana L. Lyra: Conceptualization, Re­
itat preference) and morphological traits. Alternatively, there may have sources, Data curation, Visualization, Writing – review & editing. Diego
been insufficient time for them to expand their ranges (Hutchinson, Baldo: Data curation, Visualization, Writing – review & editing. Mar­
1959; Mayr, 1964). Considering that acoustic and external morpholog­ celo Gehara: Formal analysis, Methodology, Visualization, Writing –
ical traits are not under strong divergent selection (Magalhães et al., review & editing. Célio F.B. Haddad: Resources, Funding acquisition,
2020), and that species share dietary and reproductive site preferences Visualization, Writing – review & editing. Adrian A. Garda: Concep­
(FMM personal observation), competition would enforce allopatry tualization, Resources, Funding acquisition, Supervision, Writing – re­
because morphological and/or behavioral (for frog species; Blair, 1964) view & editing.
divergence is required before two taxa can coexist (Hutchinson, 1959;
Wang and Summers, 2010). Conversely, the lack of phenotypic differ­
entiation may imply that other intrinsic sources for ecological isolation, Declaration of Competing Interest
such as specific physiological adaptations (Lexer and Fay, 2005), are
constraining species to coexist in divergent habitats (Hutchinson, 1959; The authors declare that they have no known competing financial
Zink, 2014), therefore weakening selection towards morphological and interests or personal relationships that could have appeared to influence
behavioral differentiation. Moreover, a hypothesis of recent diversifi­ the work reported in this paper.
cation is unlikely given that the last speciation event between species in
the L. latrans complex is old, dating back to late Miocene and that our Acknowledgments
results (ENMs and demography) support populational range expansion
since LIG. This work would never be possible without the help and collabora­
On the other hand, high levels of sympatry between species in the tion of several researches. FMM thanks the staff of Herpetology Labo­
L. latrans complex and a sister taxon, L. macrosternum, were observed ratory and Zoology Museums from LGE (UNaM), UFBA, UFMS, UFPB,
throughout several localities across Brazil and Argentina (Magalhães UNB, and UNESP-Rio Claro for their kind assistance and for all logistic
et al., 2020). These species also share preferences for reproductive sites support. FMM is indebted to E.F. Oliveira, E.M. Fonseca, F.M. Lanna, V.
and a generalist feeding habits, making them potential competitors for A. São-Pedro, and W. Pessoa for their support during field work. FMM
spatial and energetic resources. Therefore, one could hypothesize that thanks CNPq (Process number 167148) for his doctoral fellowship and
L. macrosternum is an additional biotic factor that could reinforce an Instituto Chico Mendes de Conservação da Biodiversidade (ICMBio,
allopatric isolation among species in the L. latrans complex. Neverthe­ Process number 53175-1) for scientific collection permits. FC thanks
less, these hypotheses should be tested in future contributions under CNPq for his postdoctoral fellowship (Process number 302162/2020-8).
comparative and predictive phylogeography frameworks (Zamudio CFBH thanks São Paulo Research Foundation (FAPESP) for financial
et al., 2016) in order to assess the intrinsic variation of physiological support (proc. #2013/50741-7) and CNPq for a research fellowship
traits and understand the mechanisms behind phenotypic selection (306623/2018-8). This research was supported by resources supplied by
within species in the L. latrans complex relative to adaptations along the High-Performance Computing Center (NPAD) at Universidade Fed­
heterogeneous environments. eral do Rio Grande do Norte-Brazil and Cyberinfrastructure for Phylo­
genetic Research (CIPRES).
5. Final remarks
Appendix A. Supplementary data
The scenario presented herein corroborates the existence of lineages
that are phylogenetically structured at both species and populational Supplementary data to this article can be found online at https://doi.
levels (Fig. 3). Interestingly, all divergent lineages are geographically org/10.1016/j.ympev.2022.107398.
partitioned but did not exhibit any major topographic barrier that would
enforce allopatry (except for L. paranaru). Instead, we found evidence
References
that most interspecific genetic breaks coincide with areas of environ­
mental transition with a second pulse of synchronous intraspecific Ab’Saber, A.N., 1977. Domínios morfoclimáticos na América do Sul: primeira
diversification during late Pleistocene, showing that climatic oscillations aproximação. Geomorfologia 52, 1–21.
played a major role in shaping genetic diversity among the Leptodactylus Auler, A.S., Smart, P.L., 2001. Late Quaternary paleoclimate in semiarid northeastern
Brazil from U-Series dating of travertine and water-table speleothems. Quat. Res. 55
latrans species complex, highlighting a congruent pattern on the primary (2), 159–167. https://doi.org/10.1006/qres.2000.2213.
mode of genetic diversification in taxa adapted to different ecoregions in Auler, A.S., Wang, X., Edwards, R.L., Cheng, H., Cristalli, P.S., Smart, P.L., Richards, D.
South America. Future works that explicitly test for contrasting models A., 2004. Quaternary ecological and geomorphic changes associated with rainfall
events in presently semi-arid northeastern Brazil. J. Quat. Sci. 19 (7), 693–701.
of isolation mechanisms under coalescent simulations will certainly https://doi.org/10.1002/jqs.876.
shed light on the evolutionary processes underlying the cryptic Avise, J.C., 2000. Phylogeography: the history and formation of species. Harvard
environment-driven diversification patterns revealed for this wide­ University Press, Cambridge, Massachusetts.
Barros, M.J.F., Silva-Arias, G.A., Fregonezi, J.N., Turchetto-Zolet, A.C., Iganci, J.R.V.,
spread frog group. Diniz-Filho, J.A.F., Freitas, L.B., 2015. Environmental drivers of diversity in
subtropical highland grasslands. Perspect. Plant Ecol., Evol. Systemat. 17 (5),
Funding 360–368. https://doi.org/10.1016/j.ppees.2015.08.001.
Batalha-Filho, H., Waldschmidt, A.M., Campos, L.A.O., Tavares, M.G., Fernandes-
Salomão, T.M., 2010. Phylogeography and historical demography of the neotropical
This study was funded by Brazilian Research Council (CNPq; Process stingless bee Melipona quadrifasciata (Hymenoptera, Apidae): incongruence between
number #140649/2015-8 to FMM; and CNPq/ICMBio #552031/2011-9 morphology and mitochondrial DNA. Apidologie 41 (5), 534–547. https://doi.org/
10.1051/apido/2010001.
to AAG) and the São Paulo Research Foundation (FAPESP; #2013/
Behling, H., 1995. Investigations into the Late Pleistocene and Holocene history of
50741-7 to CFBH). vegetation and climate in Santa Catarina (S Brazil). Vegetat. History Archaeobot. 4,
127–152. https://www.jstor.org/stable/23417538.
Behling, H., 2002. South and southeast Brazilian grasslands during Late Quaternary
CRediT authorship contribution statement
times: a synthesis. Palaeogeogr. Palaeoclimatol. Palaeoecol. 177 (1-2), 19–27.
https://doi.org/10.1016/S0031-0182(01)00349-2.
Felipe de M. Magalhães: Conceptualization, Methodology, Soft­ Behling, H., Lichte, M., 1997. Evidence of dry and cold climatic conditions at glacial
ware, Formal analysis, Investigation, Data curation, Writing – original times in tropical Southeastern Brazil. Quat. Res. 48 (3), 348–358. https://doi.org/
10.1006/qres.1997.1932.
draft, Writing – review & editing. Felipe Camurugi: Conceptualization, Blair, W.F., 1964. Isolating mechanisms and interspecies interactions in anuran
Data curation, Formal analysis, Software, Visualization, Methodology, amphibians. Q. Rev. Biol. 39 (4), 334–344. https://doi.org/10.1086/404324.

12
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

Bocalini, F., Bolívar-Leguizamón, S.D., Silveira, L.F., Bravo, G.A., 2021. Comparative Firkowski, C.R., Bornschein, M.R., Ribeiro, L.F., Pie, M.R., 2016. Species delimitation,
phylogeographic and demographic analyses reveal a congruent pattern of sister phylogeny and evolutionary demography of co-distributed, montane frogs in the
relationships between bird populations of the northern and south-central Atlantic southern Brazilian Atlantic Forest. Mol. Phylogenet. Evol. 100, 345–360. https://doi.
Forest. Mol. Phylogenet. Evol. 154, 106973. https://doi.org/10.1016/j. org/10.1016/j.ympev.2016.04.023.
ympev.2020.106973. Fitzpatrick, S.W., Brasileiro, C.A., Haddad, C.F.B., Zamudio, K.R., 2009. Geographical
Bouckaert, R.R., 2010. DensiTree: making sense of sets of phylogenetic trees. variation in genetic structure of an Atlantic Coastal Forest frog reveals regional
Bioinformatics 26 (10), 1372–1373. https://doi.org/10.1093/bioinformatics/ differences in habitat stability. Mol. Ecol. 18, 2877–2896. https://doi.org/10.1111/
btq110. j.1365-294X.2009.04245.x.
Bouckaert, R., Vaughan, T.G., Barido-Sottani, J., Duchêne, S., Fourment, M., Flouri, T., Jiao, X., Rannala, B., Yang, Z., 2018. Species tree inference with BPP using
Gavryushkina, A., Heled, J., Jones, G., Kühnert, D., Maio, N., Matschiner, M., genomic sequences and the multispecies coalescent. Mol. Biol. Evol. 35 (10),
Mendes, F.K., Müller, N.F., Ogilvie, H.A., Plessis, L., Popinga, A., Rambaut, A., 2585–2593. https://doi.org/10.1093/molbev/msy147.
Rasmussen, D., Siveroni, I., Suchard, M.A., Wu, C.H., Xie, D., Zhang, C., Stadler, T., Fonte, L.M., Mayer, M., Lötters, S., 2019. Long-distance dispersal in amphibians. Front.
Drummond, A.J., 2019. BEAST 2.5: an advanced software platform for Bayesian Biogeogr. 11 (4) https://doi.org/10.21425/F5FBG44577.
evolutionary analysis. PLoS Comput. Biol. 15 (4), e1006650. https://doi.org/ Fouquet, A., Noonan, B.P., Rodrigues, M.T., Pech, N., Gilles, A., Gemmell, N.J., 2012.
10.1371/journal.pcbi.1006650. Multiple quaternary refugia in the eastern Guiana Shield revealed by comparative
Broennimann, O., Fitzpatrick, M.C., Pearman, P.B., Petitpierre, B., Pellissier, L., phylogeography of 12 frog species. Syst. Biol. 61, 461–489. https://doi.org/
Yoccoz, N.G., Thuiller, W., Fortin, M.J., Randin, C., Zimmermann, N.E., Graham, C. 10.1093/sysbio/syr130.
H., Guisan, A., 2012. Measuring ecological niche overlap from occurrence and Fu, Y.X., 1997. Statistical tests of neutrality of mutations against population growth,
spatial environmental data. Glob. Ecol. Biogeogr. 21, 481–497. https://doi.org/ hitchhiking and background selection. Genetis 147 (2), 915–925.
10.1111/j.1466-8238.2011.00698.x. Gehara, M., Barth, A., Oliveira, E.F., Costa, M.A., Haddad, C.F.B., Vences, M., 2017a.
Bruford, M.W., Hanotte, O., Brookfield, J.F.Y., Burke, T., 1992. Single and multilocus Model-based analyses reveal insular population diversification and cryptic frog
DNA fingerprinting. In: Hoelzel, A.R. (Ed.), Molecular Genetic Analysis of species in the Ischnocnema parva complex in the Atlantic forest of Brazil. Mol.
Populations: A Practical Approach. Press, Oxford, I. R. L, pp. 225–269. Phylogenet. Evol. 112, 68–78. https://doi.org/10.1016/j.ympev.2017.04.007.
Bruschi, D.P., Peres, E.A., Lourenço, L.B., Bartoleti, L.F.M., Sobral-Souza, T., Recco- Gehara, M., Garda, A.A., Werneck, F.P., Oliveira, E.F., Fonseca, E.M., Camurugi, F.,
Pimentel, S.M., 2019. Signature of the paleo-course changes in the São Francisco Magalhães, F.M., Lanna, F.M., Sites Jr., J.W., Marques, R., Silveira-Filho, R., São-
river as source of genetic structure in Neotropical Pithecopus nordestinus Pedro, V.A., Colli, G.R., Costa, G.C., Burbrink, F.T., 2017b. Estimating synchronous
(Phyllomedusinae, Anura) treefrog. Front. Genet. 10 https://doi.org/10.3389/ demographic changes across populations using hABC and its application for a
fgene.2019.00728. herpetological community from northeastern Brazil. Mol. Ecol. 18, 4756–4771.
Byrne, M., 2008. Evidence for multiple refugia at different time scales during Pleistocene https://doi.org/10.1111/mec.14239.
climatic oscillations in southern Australia inferred from phylogeography. Quat. Sci. Giraudo, A.R., Arzamendia, V., 2017. Descriptive bioregionalisation and conservation
Rev. 27 (27-28), 2576–2585. https://doi.org/10.1016/j.quascirev.2008.08.032. biogeography: what is the true bioregional representativeness of protected areas?
Cabanne, G.S., Calderón, L., Arias, N.T., Flores, P., Pessoas, R., d’Horta, F.M., Miyaki, C. Aust. Syst. Bot. 30, 403–413. https://doi.org/10.1071/SB16056.
Y., 2016. Effects of Pleistocene climate changes on species ranges and evolutionary Grazziotin, F.G., Monzel, M., Echeverrigaray, S., Bonatto, S.L., 2006. Phylogeography of
processes in the Neotropical Atlantic Forest. Biol. J. Linn. Soc. 119 (4), 856–872. the Bothrops jararaca complex (Serpentes: Viperidae): past fragmentation and island
https://doi.org/10.1111/bij.12844. colonization in the Brazilian Atlantic Forest. Mol. Ecol. 15, 3969–3982. https://doi.
Camurugi, F., Gehara, M., Fonseca, E.M., Zamudio, K.R., Haddad, C.F.B., Colli, G.R., org/10.1111/j.1365-294X.2006.03057.x.
Thomé, M.T., Prado, C.P.A., Napoli, M.F., Garda, A.A., 2021. Isolation by Guillot, G., Mortier, F., Estoup, A., 2005. Geneland: a computer package for landscape
environment and recurrent gene flow shaped the evolutionary history of a genetics. Mol. Ecol. Notes 5 (3), 712–715. https://doi.org/10.1111/j.1471-
continentally distributed Neotropical treefrog. J. Biogeogr. 48 (4), 760–772. https:// 8286.2005.01031.x.
doi.org/10.1111/jbi.14035. Hansen, J., Sato, M., Russell, G., Kharecha, P., 2013. Climate sensitivity, sea level and
Carnaval, A.C., Hickerson, M.J., Haddad, C.F.B., Rodrigues, M.T., Moritz, C., 2009. atmospheric carbon dioxide. Philos. Trans. Royal Soc. A 371 (2001), 20120294.
Stability predicts genetic diversity in the Brazilian Atlantic Forest hotspot. Science https://doi.org/10.1098/rsta.2012.0294.
323 (5915), 785–789. Hasumi, H., Emori, S., 2004. K-1 coupled GCM (MIROC) description. University of
Carnaval, A.C., Moritz, C., 2008. Historical climate modelling predicts patterns of current Tokyo, Tokyo, Japan, Center for Climate System Research.
biodiversity in the Brazilian Atlantic forest. J. Biogeogr. 35, 1187–1201. https://doi. Heled, J., Drummond, A.J, 2008. Bayesian inference of population size history from
org/10.1111/j.1365-2699.2007.01870.x. multiple loci. BMC Evol. Biol. 8 (1), 289. https://doi.org/10.1186/1471-2148-8-
Carnaval, A.C., Waltari, E., Rodrigues, M.T., Rosauer, D., VanDerWal, J., Damasceno, R., 289.
Prates, I., Strangas, M., Spanos, Z., Rivera, D., Pie, M.R., Firkowski, C.R., Hijmans, R.J., Cameron, S.E., Parra, J.L., Jones, P.G., Jarvis, A., 2005. Very high
Bornschein, M.R., Ribeiro, L.F., Moritz, C., 2014. Prediction of phylogeographic resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25
endemism in an environmentally complex biome. Proceed. Royal Soc. B: Biol. Sci. (15), 1965–1978. https://doi.org/10.1002/joc.1276.
281 (1792), 20141461. https://doi.org/10.1098/rspb.2014.1461. Hoorn, C., Wesselingh, F.P., ter Steege, H., Bermudez, M.A., Mora, A., Sevink, J.,
Carstens, B., Richards, C.L., 2007. Integrating coalescent and ecological niche modeling Sanmartín, I., Sanchez-Meseguer, A., Anderson, C.L., Figueiredo, J.P., Jaramillo, C.,
in comparative phylogeography. Evolution 61 (6), 1439–1454. https://doi.org/ Riff, D., Negri, F.R., Hooghiemstra, H., Lundberg, J., Stadler, T., Särkinen, T.,
10.1111/j.1558-5646.2007.00117.x. Antonelli, A., 2010. Amazonia through time: Andean uplift, climate change,
Colombi, V.H., Lopes, S.R., Fagundes, V., 2010. Testing the Rio Doce as a riverine barrier landscape evolution, and biodiversity. Science 330 (6006), 927–931.
in shaping the Atlantic rainforest population divergence in the rodent Akodon cursor. Hutchinson, G.E., 1959. Homage to Santa Rosalia or why are there so many kinds of
Genet. Mol. Biol. 33 (4), 785–789. https://doi.org/10.1590/S1415- animals? Am. Nat. 93 (870), 145–159. https://doi.org/10.1086/282070.
47572010000400029. Irwin, D.E., 2002. Phylogeographic breaks without geographic barriers to gene flow.
Costa, L.P., Leite, Y.L.R., 2012. Historical fragmentation shaping vertebrate Evolution 56 (12), 2383–2394. https://doi.org/10.1111/j.0014-3820.2002.
diversification in the Atlantic Forest biodiversity hotspot. In: Patterson, P., Costa, L. tb00164.x.
P. (Eds.), Bones, clones, and biomes: the history and geography of recent Neotropical Janes, J.K., Miller, J.M., Dupuis, J.R., Malenfant, R.M., Gorrel, J.C., Cullingham, C.I.,
mammals. The University of Chicago Press, pp. 283–306. Andrew, R.L., 2017. The K = 2 conundrum. Mol. Ecol. 26 (14), 3594–3602. https://
d’Horta, F.M., Cabanne, G.S., Meyer, D., Miyaki, C.Y., 2011. The genetic effects of Late doi.org/10.1111/mec.14187.
Quaternary climatic changes over a tropical latitudinal gradient: diversification of an Katoh, K., Misawa, K., Kuma, K., Miyata, T., 2002. MAFFT: a novel method for rapid
Atlantic Forest passerine. Mol. Ecol. 20, 1923–1935. https://doi.org/10.1111/ multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 30,
j.1365-294X.2011.05063.x. 3059–3066. https://doi.org/10.1093/nar/gkf436.
Darriba, D., Taboada, G.L., Doallo, R., Posada, D., 2012. jModelTest 2: more models, new Kearse, M., Moir, R., Wilson, A., Stones-Havas, S., Cheung, M., Sturrock, S., Buxton, S.,
heuristics and parallel computing. Nat. Methods 9 (8), 772. https://doi.org/ Cooper, A., Markowitz, S., Duran, C., Thierer, T., Ashton, B., Meintjes, P.,
10.1038/nmeth.2109. Drummond, A., 2012. Geneious basic: an integrated and extendable desktop
Degnan, J.H., Rosenberg, N.A., 2009. Gene tree discordance, phylogenetic inference and software platform for the organization and analysis of sequence data. Bioinformatics
the multispecies coalescent. Trends Ecol. Evol. 24 (6), 332–340. https://doi.org/ 28 (12), 1647–1649. https://doi.org/10.1093/bioinformatics/bts199.
10.1016/j.tree.2009.01.009. Kopuchian, C., Campagna, L., Lijtmaer, D.A., Cabanne, G.S., García, N.C., Lavinia, P.D.,
Di Cola, V., Broennimann, O., Petitpierre, B., Breiner, F.T., D’Amen, M., Randin, C., Tubaro, P.L., Lovette, I., Di-Giacomo, A.S., 2020. A test of the riverine barrier
Engler, R., Pottier, J., Pio, D., Dubuis, A., Pellissier, L., Mateo, R.G., Hordijk, W., hypothesis in the largest subtropical river basin in the Neotropics. Mol. Ecol. 29 (12),
Salamin, N., Guisan, A., 2017. Ecospat: an R package to support spatial analyses and 2137–2149. https://doi.org/10.1111/mec.15384.
modeling of species niches and distributions. Ecography 40 (6), 774–787. https:// Lanna, F.M., Gehara, M., Werneck, F.P., Fonseca, E.M., Colli, G.R., Sites Jr., J.W.,
doi.org/10.1111/ecog.02671. Rodrigues, M.T., Garda, A.A., 2020. Dwarf geckos and giant rivers: the role of the
Earl, D.A., vonHoldt, B.M., 2012. STRUCTURE HARVESTER: a website and program for São Francisco River in the evolution of Lygodactylus klugei (Squamata: Gekkonidae)
visualizing STRUCTURE output and implementing the Evanno method. Conserv. in the semi-arid Caatinga of north-eastern Brazil. Biol. J. Linn. Soc. 129, 88–98.
Genet. Resour. 4, 359–361. https://doi.org/10.1007/s12686-0119548-7. https://doi.org/10.1093/biolinnean/blz170.
Evanno, G., Regnaut, S., Goudet, J., 2005. Detecting the number of clusters of individuals Leaché, A.D., Fujita, M.K., 2010. Bayesian species delimitation in West African forest
using the software STRUCTURE: a simulation study. Mol. Ecol. 14 (8), 2611–2620. geckos (Hemidactylus fasciatus). Proceedings of the Royal Society B 277 (1697),
https://doi.org/10.1111/j.1365-294X.2005.02553.x. 3071–3077. https://doi.org/10.1098/rspb.2010.0662.
Fazolato, C., Fernandes, F., Batalha-Filho, H., 2017. The effects of Quaternary sea-level Leite, Y.L.R., Costa, L.P., Loss, A.C., Rocha, R.G., Batalha-Filho, H., Bastos, A.C.,
fluctuations on the evolutionary history of an endemic ground lizard (Tropidurus Quaresma, V.S., Fagundes, V., Paresque, R., Passamani, M., Pardini, R., 2016.
hygomi). Zool. Anz. 270, 1–8. https://doi.org/10.1016/j.jcz.2017.08.007. Neotropical forest expansion during the last glacial period challenges refuge

13
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

hypothesis. Proc. NatL. Acad. Sci. USA 113 (4), 1008–1013. https://doi.org/ Peçanha, W.T., Althoff, S.L., Galiano, D., Quintela, F.M., Maestri, R., Gonçalves, G.L.,
10.1073/pnas.1513062113. Freitas, T.R.O., 2017. Pleistocene climatic oscillations in Neotropical open areas:
Lexer, C., Fay, M.F., 2005. Adaptation to environmental stress: a rare or frequent driver refuge isolation in the rodent Oxymycterus nasutus endemic to grasslands. PLoS ONE
of speciation? J. Evol. Biol. 18 (4), 893–900. 12 (11), e0187329. https://doi.org/10.1371/journal.pone.0187329.
Librado, P., Rozas, J., 2009. DnaSP v5: A software for comprehensive analysis of DNA Phillips, S.J., Anderson, R.P., Schapire, R.E., 2006. Maximum entropy modeling of
polymorphism data. Bioinformatics 25 (11), 1451–1452. https://doi.org/10.1093/ species geographic distributions. Ecol. Model. 190 (3-4), 231–259. https://doi.org/
bioinformatics/btp187. 10.1016/j.ecolmodel.2005.03.026.
Lima, C.C.U., Nolasco, M.C., 2015. Chapada Diamantina: a remarkable landscape Popolizio, E., 2004. El Paraná, un río y su historia geomorfológica. Tomos I y II. (PhD
dominated by mountains and plateaus, in: Vieira, B.C., Salgado, A.A.R., Santos, L.J. thesis). Centro de Geociencias Aplicadas. Resistencia, Tomo 19, 1–362.
C. (Eds.). Landscapes and Landforms of Brazil. World Geomorphological Landscapes. Pritchard, J., Stephens, M., Donnelly, P., 2000. Inference of population structure using
Springer, Dordrecht, pp. 211–220. http://doi.org/10.1007/978-94-017-8023-0_19. multilocus genotype data. Genetics 155 (2), 945–959.
Lynch, M., 2007. The origins of genome architecture. Sinauer Associates, Sunderland. Proosdij, A.S., Sosef, M.S.M., Wieringa, J.J., Raes, N., 2016. Minimum required number
Lyra, M.L., Haddad, C.F.B., de Azeredo-Espin, A.M.L., 2017. Meeting the challenge of of specimen records to develop accurate species distribution models. Ecography 39
DNA barcoding Neotropical amphibians: polymerase chain reaction optimization (6), 542–552. https://doi.org/10.1111/ecog.01509.
and new COI primers. Mol. Ecol. Resour. 17 (5), 966–980. https://doi.org/10.1111/ Pyron, R.A., Burbrink, F.T., 2009. Lineage diversification in a widespread species: roles
1755-0998.12648. for niche divergence and conservatism in the common kingsnake, Lampropeltis getula.
Mabesoone, J.M., 1994. Sedimentary basins of Northeast Brazil, 1ª ed. Editora Mol. Ecol. 18, 3443–3457. https://doi.org/10.1111/j.1365-294X.2009.04292.x.
Universitária, Universidade Federal de Pernambuco, Recife. Pyron, R.A., Burbrink, F.T., 2010. Hard and soft allopatry: physically and ecologically
Magalhães, F.M., Lyra, L.M., de Carvalho, T.R., Baldo, D., Brusquetti, F., Burella, P., mediated modes of geographic speciation. J. Biogeogr. 37, 2005–2015. https://doi.
Colli, G.R., Gehara, M.C., Giaretta, A.A., Haddad, C.F.B., Langone, J.A., López, J.A., org/10.1111/j.1365-2699.2010.02336.x.
Napoli, M.F., Santana, D.J., de Sá, R.O., Garda, A.A., 2020. Taxonomic review of R Core Team. 2018. R: A language and environment for statistical computing. Vienna, R
South American butter frogs: phylogeny, geographic patterns, and species Foundation for Statistical Computing, Austria. Available at: http://www.R-project.or
delimitation in the Leptodactylus latrans species group (Anura: Leptodactylidae). g.
Herpetol. Monographs 34, 131–177. https://doi.org/10.1655/HERPMONOGRAPHS- Rambaut, A., 2014. FigTree: a graphical viewer of phylogenetic trees, version 1.4.2.
D-19-00012. Available at: http://tree.bio.ed.ac.uk/software/figtree/.
Maia-Carvalho, B., Vale, C.G., Sequeira, F., Ferrand, N., Martínez-Solano, I., Rambaut, A., Drummond, A.J., Xie, D., Baele, G., Suchard, M.A., 2018. Posterior
Gonçalves, H., 2018. The roles of allopatric fragmentation and niche divergence in summarization in Bayesian phylogenetics using Tracer 1.7. Syst. Biol. 67, 901–904.
intraspecific lineage diversification in the common midwife toad (Alytes obstetricans). https://doi.org/10.1093/sysbio/syy032.
J. Biogeogr. 45 (9), 2146–2158. https://doi.org/10.1111/jbi.13405. Ramos-Onsins, S.E., Rozas, J., 2002. Statistical properties of new neutrality tests against
Maniatis, T., Fritsch, E.F., Sambrook, J., 1982. Molecular Cloning: A Laboratory Manual. population growth. Mol. Biol. Evol. 19, 2092–2100. https://doi.org/10.1093/
Cold Spring Harbour Laboratory, USA. oxfordjournals.molbev.a004034.
Márquez, A., Maldonado, J.E., González, S., Beccaceci, M.D., Garcia, J.E., Duarte, J.M.B., Ribeiro, A.C., Lima, F.C.T., Riccomini, C., Menezes, N.A., 2006. Fishes of the Atlantic
2006. Phylogeography and Pleistocene demographic history of the endangered Rainforest of Boracéia: testimonies of the Quaternary fault reactivation within a
marsh deer (Blastocerus dichotomus) from the Río de la Plata Basin. Conserv. Genet. 7, Neoproterozoic tectonic province in Southeastern Brazil. Ichthyol. Explor.
563–575. https://doi.org/10.1007/s10592-005-9067-8. Freshwaters 17, 157–164.
Martins, F.M., 2011. Historical biogeography of the Brazilian Atlantic Forest and the Rissler, L.J., Apodaca, J.J., Weins, J., 2007. Adding more ecology into species
Carnaval-Moritz model of Pleistocene refugia: what do phylogeographical studies delimitation: ecological niche models and phylogeography help define cryptic
tell us? Biol. J. Linn. Soc. 104, 499–509. https://doi.org/10.1111/j.1095- species in the black salamander (Aneides flavipunctatus). Syst. Biol. 56 (6), 924–942.
8312.2011.01745.x. https://doi.org/10.1080/10635150701703063.
Mayr, E., 1947. Ecological factors in speciation. Evolution 1, 263–288. https://doi.org/ Rödder, D., Engler, J.O., 2011. Quantitative metrics of overlaps in Grinnellian niches:
10.2307/2405327. advances and possible drawbacks. Glob. Ecol. Biogeogr. 20, 915–927. https://doi.
Inger, Robert F., Mayr, Ernst, 1964. Animal species and evolution. Copeia 1964 (1), 245. org/10.1111/j.1466-8238.2011.00659.x.
https://doi.org/10.2307/1440881. Rodríguez, A., Borner, M., Pabijan, M., Gehara, M., Haddad, C.F.B., Vences, M., 2015.
Mello, C.L., Metelo, C.M.S., Suguio, K., Kohler, H.C., 1999. Quaternary sedimentation, Genetic divergence in tropical anurans: deeper phylogeographic structure in forest
neotectonics and the evolution of the Doce River Middle Valley lake system specialists and in topographically complex regions. Evol. Ecol. 29 (5), 765–785.
(southeastern Brazil). Revista do Instituto Geológico, IG São Paulo 2, 29–36. https://doi.org/10.1007/s10682-015-9774-7.
Menezes, R.S.T., Brady, S.G., Carvalho, A.F., Del Lama, M.A., Costa, M.A., 2017. The Rull, V., 2008. Speciation timing and neotropical biodiversity: the Tertiary-Quaternary
roles of barriers, refugia, and chromosomal clines underlying diversification in debate in the light of molecular phylogenetic evidence. Mol. Ecol. 17, 2722–2729.
Atlantic Forest social wasps. Sci. Rep. 7, 7689. https://doi.org/10.1038/s41598- https://doi.org/10.1111/j.1365-294X.2008.03789.x.
017-07776-7. Rull, V., 2011. Neotropical biodiversity: timing and potential drivers. Trends Ecol. Evol.
Mittermeier, R.A., Myers, N., Thomsen, J.B., da Fonseca, G.A.B., Olivieri, S., 1998. 26 (10), 508–513. https://doi.org/10.1016/j.tree.2011.05.011.
Biodiversity hotspots and major tropical wilderness areas: approaches to setting Saadi, A., Machette, M.N., Haller, K.M., Dart, R.L., Bradley, L., Souza, A.M.P.D., 2002.
conservation priorities. Conserv. Biol. 12 (3), 516–520. https://doi.org/10.1046/ Map and database of Quaternary faults and lineaments in Brazil. U.S. Geological
j.1523-1739.1998.012003516.x. Survey. Open-File Report 02-230, Version 1.0. Available from: <http://pubs.usgs.
Morellato, L.P.C., Haddad, C.F.B., 2000. Introduction: the Brazilian Atlantic forest. gov/of/2002/ofr-02-230/>.
Biotropica 32, 786–792. https://doi.org/10.1111/j.1744-7429.2000.tb00618.x. Sabbag, A.F., Lyra, L.M., Zamudio, K.R., Haddad, C.F.B., Feio, R.N., Leite, F.S.F.,
Muscarella, R., Soley-Guardia, M., Boria, R.A., Kass, J.M., Uriarte, M., Anderson, R.P., Gasparini, J.L., Brasileiro, C.A., 2018. Molecular phylogeny of Neotropical rock frogs
2014. ENMeval: An R package for conducting spatially independent evaluations and reveals a long history of vicariant diversification in the Atlantic forest. Mol.
estimating optimal model complexity for Maxent ecological niche models. Methods Phylogenet. Evol. 2018, 142–156. https://doi.org/10.1016/j.ympev.2018.01.017.
Ecol. Evol. 5, 1198–1205. https://doi.org/10.1111/2041-210X.12261. Schoener, T.W., 1968. The Anolis lizards of Bimini: resource partitioning in a complex
Myers, E.A., McKelvy, A.D., Burbrink, F.T., 2020. Biogeographic barriers, Pleistocene fauna. Ecology 49, 704–726. https://doi.org/10.2307/1935534.
refugia, and climatic gradients in the southeastern Nearctic drive diversification in Senczuk, G., Harris, D.J., Castiglia, R., Litsi Mizan, V., Colangelo, P., Canestrelli, D.,
cornsnakes (Pantherophis guttatus complex). Mol. Ecol. 29, 797–811. https://doi.org/ Salvi, D., 2019. Evolutionary and demographic correlates of Pleistocene coastline
10.1111/mec.15358. changes in the Sicilian wall lizard Podarcis wagleriana. J. Biogeogr. 46 (1), 224–237.
Naimi, B., Hamm, N.A.S., Groen, T.A., Skidmore, A.K., Toxopeus, A.G., 2014. Where is https://doi.org/10.1111/jbi.13479.
positional uncertainty a problem for species distribution modelling? Ecography 37 Shafer, A.B.A., Wolf, J.B.W., Wiens, J., 2013. Widespread evidence for incipient
(2), 191–203. https://doi.org/10.1111/j.1600-0587.2013.00205.x. ecological speciation: a meta-analysis of isolation-by-ecology. Ecol. Lett. 16 (7),
Nascimento, F.F., Lazar, A., Menezes, A.N., Durans, A.M., Moreira, J.C., Salazar- 940–950. https://doi.org/10.1111/ele.12120.
Bravo, J., D′ Andrea, P.S., Bonvicino, C.R., Fine, P.V.A., 2013. The role of historical Siddal, M., Rohling, E.J., Blunier, T., Spahni, R., 2010. Patterns of millennial variability
barriers in the diversification processes in open vegetation formations during the over the last 500 ka. Clim. Past 6, 295–303. https://doi.org/10.5194/cp-6-295-
Miocene/Pliocene using an ancient rodent lineage as a model. PLoS ONE 8 (4), 2010.
e61924. https://doi.org/10.1371/journal.pone.0061924. Smith, B.T., McCormack, J.E., Cuervo, A.M., Hickerson, M.J., Aleixo, A., Cadena, C.D.,
Ogilvie, H.A., Bouckaert, R.R., Drummond, A.J., 2017. StarBEAST2 brings faster species Pérez-Emán, J., Burney, C.W., Xie, X., Harvey, M.G., Faircloth, B.C., Glenn, T.C.,
tree inference and accurate estimates of substitution rates. Mol. Biol. Evol. 34, Derryberry, E.P., Prejean, J., Fields, S., Brumfield, R.T., 2014. The drivers of tropical
2101–2114. https://doi.org/10.1093/molbev/msx126. speciation. Nature 515 (7527), 406–409. https://doi.org/10.1038/nature13687.
Olson, D.M., Dinerstein, E., Wikramanayake, E.D., Burgess, N.D., Powell, G.V.N., Sotelo-Muñoz, M., Maldonado-Coelho, M., Svensson-Coelho, M., dos Santos, S.S.,
Underwood, E.C., D’amico, J.A., Itoua, I., Strand, H.E., Morrison, J.C., Loucks, C.J., Miyaki, C.Y., 2020. Vicariance, dispersal, extinction and hybridization underlie the
Allnutt, T.F., Ricketts, T.H., Kura, Y., Lamoreux, J.F., Wettengel, W.W., Hedao, P., evolutionary history of Atlantic forest fire-eye antbirds (Aves: Thamnophilidae).
Kassem, K.R., 2001. Terrestrial ecoregions of the world: a new map of life on earth: A Mol. Phylogenet. Evol. 148, 106820. https://doi.org/10.1016/j.
new global map of terrestrial ecoregions provides an innovative tool for conserving ympev.2020.106820.
biodiversity. Bioscience 51, 933–938. https://doi.org/10.1641/0006-3568(2001) Souza, C.R.G., 2015. The Bertioga coastal plain: an example of morphotectonic
051[0933:TEOTWA]2.0.CO;2. evolution, in: Vieira, B.C., Salgado, A.A.R., Santos, L.J.C. (Eds.). Landscapes and
Otto-Bliesner, B.L., Marshall, S.J., Overpeck, J.T., Miller, G.H., Hu, A., 2006. CAPE Last Landforms of Brazil. World Geomorphological Landscapes. Springer, Dordrecht, pp.
Interglacial Project members: Simulating arctic climate warmth and icefield retreat 115–134. http://doi.org/10.1007/978-94-017-8023-0_11.
in the last interglaciation. Science 311 (5768), 1751–1753.

14
F.M. Magalhães et al. Molecular Phylogenetics and Evolution 169 (2022) 107398

Stephens, M., Smith, N.J., Donnelly, P., 2001. A new statistical method for haplotype Vuilleumier, B.S., 1971. Pleistocene changes in the fauna and flora of South America.
reconstruction from population data. Am. J. Hum. Genet. 68 (4), 978–989. https:// Science 173 (3999), 771–780.
doi.org/10.1086/319501. Wang, I.J., Bradburd, G.S., 2014. Isolation by environment. Mol. Ecol. 23 (23),
Suguio, K., Martins, L., Bittencourt, A.C.S.P., Dominguez, J.M.L., Flexor, J., Azevedo, A. 5649–5662. https://doi.org/10.1111/mec.12938.
E.G., 1985. Flutuações do nível relativo do mar durante o Quaternário superior ao Wang, I.J., Glor, R.E., Losos, J.B., 2013. Quantifying the roles of ecology and geography
longo do litoral brasileiro e suas implicações na sedimentação costeira. Revista in spatial genetic divergence. Ecol. Lett. 16 (2), 175–182. https://doi.org/10.1111/
Brasileira de Geociências 15, 273–286. ele.12025.
Tajima, F, 1989. Statistical method for testing the neutral mutation hypothesis by DNA Wang, I.J., Summers, K., 2010. Genetic structure is correlated with phenotypic
polymorphism. Genetis 123 (3), 585–595. divergence rather than geographic isolation in the highly polymorphic strawberry
Thomé, M.T., Zamudio, K.R., Giovanelli, J.G., Haddad, C.F.B., Baldissera, F.A., poison-dart frog. Mol. Ecol. 19 (3), 447–458.
Alexandrino, J., 2010. Phylogeography of endemic toads and post-Pliocene Wang, X., Auler, A.S., Edwards, R.L., Cheng, H., Cristalli, P.S., Smart, P.L., Richards, D.
persistence of the Brazilian Atlantic Forest. Mol. Phylogenet. Evol., 55, 1018–1031. A., Shen, C., 2004. Wet periods in northeastern Brazil over the past 210 kyr linked to
Thomé, M.T., Zamudio, K.R., Haddad, C.F.B., Alexandrino, J., 2014. Barriers, rather than distant climate anomalies. Nature 432, 740–743. https://doi.org/10.1038/
refugia, underlie the origin of diversity in toads endemic to the Brazilian Atlantic nature03067.
Forest. Mol. Ecol. 23 (24), 6152–6164. Warren, D.L., Glor, R.E., Turelli, M., 2008. Environmental niche equivalency versus
Thomé, M.T., Carstens, B.C., Rodrigues, M.T., Alexandrino, J., Haddad, C.F.B., 2021. conservatism: quantitative approaches to niche evolution. Evolution 62, 2868–2883.
Genomic data from the Brazilian sibilator frog reveal contrasting Pleistocene https://doi.org/10.1111/j.15585646.2008.00482.x.
dynamics and regionalism in two South American dry biomes. J. Biogeogr. 48 (5), Waters, J.M., 2011. Competitive exclusion: phylogeography’s ‘elephant in the room?
1112–1123. https://doi.org/10.1111/jbi.14064. Mol. Ecol. 20, 4388–4394. https://doi.org/10.1111/j.1365-294X.2011.05286.x.
Thorpe, R.S., Surget-Groba, Y., Johansson, H., 2008. The relative importance of ecology Wells, K.D., 2007. The ecology and behavior of amphibians. The University of Chicago
and geographic isolation for speciation in anoles. Philos. Trans. Royal Soc. B: Biol. Press, USA.
Sci. 363 (1506), 3071–3081. https://doi.org/10.1098/rstb.2008.0077. Werle, E., Schneider, C., Renner, M., Völker, M., Fiehn, W., 1994. Convenient single-step,
Tonini, J.F.R., Costa, L.P., Carnaval, A.C., 2013. Phylogeographic structure is strong in one tube purification of PCR products for direct sequencing. Nucleic Acids Res. 22
the Atlantic Forest; predictive power of correlative paleodistribution models, not (20), 4354–4355.
always. J. Zool. Syst. Evol. Res. 51 (2), 114–121. https://doi.org/10.1111/ Yang, Z., Rannala, B., 2014. Unguided species delimitation using DNA sequence data
jzs.12014. from multiple loci. Mol. Biol. Evol. 31 (12), 3125–3135. https://doi.org/10.1093/
Turchetto-Zolet, A.C., Pinheiro, F., Salgueiro, F., Palma-Silva, C., 2013. molbev/msu279.
Phylogeographical patterns shed light on evolutionary process in South America. Zamudio, K.R., Bell, R.C., Mason, N.A., 2016. Phenotypes in phylogeography: species’
Mol. Ecol. 22 (5), 1193–1213. https://doi.org/10.1111/mec.12164. traits, environmental variation, and vertebrate diversification. Proc. Natl. Acad. Sci.
Vasconcellos, M.M., Colli, G.R., Weber, J.N., Ortiz, E.M., Rodrigues, M.T., Cannatella, D. USA 113 (29), 8041–8048. https://doi.org/10.1073/pnas.1602237113.
C., 2019. Isolation by instability: historical climate change shapes population Zhang, C., Rannala, B., Yan, Z., 2014. Bayesian species delimitation can be robust to
structure and genomic divergence of treefrogs in the Neotropical Cerrado savanna. guide-tree inference errors. Syst. Biol. 63, 993–1004. https://doi.org/10.1093/
Mol. Ecol. 28 (7), 1748–1764. https://doi.org/10.1111/mec.15045. sysbio/syu052.
Vences, M., Thomas, M., Bonett, R.M., Vieites, D.R., 2005. Deciphering amphibian Zink, R.M., Manne, L., 2014. Homage to Hutchinson, and the role of ecology in lineage
diversity through DNA barcoding: chances and challenges. Philos. Trans. Royal Soc. divergence and speciation. J. Biogeogr. 41 (5), 999–1006. https://doi.org/10.1111/
B: Biol. Sci. 360 (1462), 1859–1868. https://doi.org/10.1098/rstb.2005.1717. jbi.12252.
Vieira, B.C., Salgado, A.A.R., Santos, L.J.C., 2015. Landscapes and Landforms of Brazil.
World Geomorphological Landscapes. Springer, Dordrecht. http://doi.org/10.1007/
978-94-017-8023-0.

15

You might also like