Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Molecular Phylogenetics and Evolution 190 (2024) 107959

Contents lists available at ScienceDirect

Molecular Phylogenetics and Evolution


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

Integrative species delimitation uncovers hidden diversity within the


Pithecopus hypochondrialis species complex (Hylidae, Phyllomedusinae) and
its phylogeography reveals Plio-Pleistocene connectivity among
Neotropical savannas
Rafael F. Magalhães a, b, *, Elisa K. S. Ramos c, Lucas N. Bandeira d, Johnny S. Ferreira f,
Fernanda P. Werneck d, e, Marina Anciães d, e, Daniel P. Bruschi f, *
a
Department of Natural Sciences, Universidade Federal de São João del-Rei, Campus Dom Bosco, Praça Dom Helvécio, 70, São João del-Rei, MG 36301-160, Brazil
b
Postgraduate Programme in Zoology, Institute of Biological Sciences, Universidade Federal de Minas Gerais, Avenida Antônio Carlos, 6627, Belo Horizonte, MG 31270-
010, Brazil
c
Faculty of Philosophy and Natural Sciences, Department of Environmental Sciences, University of Basel, Bernoullistrasse 30, Basel 4056, Switzerland
d
Postgraduate Programme in Ecology, Instituto Nacional de Pesquisas da Amazônia, Avenida André Araújo, 2936, Manaus, AM 69067-375, Brazil
e
Scientific Biological Collections Program, Biodiversity Coordination, Instituto Nacional de Pesquisas da Amazônia, Avenida André Araújo, 2936, Manaus, AM 69067-
375, Brazil
f
Postgraduate Programme in Genetics, Department of Genetics, Biological Sciences Sector, Universidade Federal do Paraná, Caixa Postal 19071, Curitiba, PR 81531-
980, Brazil

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

Keywords: Despite their limited vagility and pronounced habitat heterogeneity in the tropics, many anuran species have
Amazonia unexpectedly extensive geographic ranges. One prominent example of this phenomenon is Pithecopus hypo­
Cerrado chondrialis, which is found in the Cerrado, Guianan savanna, and Llanos domains, as well as isolated tracts of
Ecological niche modelling
savanna and open habitat within the Amazon Forest. The present study employs an integrative species delimi­
Llanos
tation approach to test the hypothesis that P. hypochondrialis is in fact a species complex. We also reconstruct the
Guianan savannas
Phylogeographic diffusion relationships among the lineages delimited here and other Pithecopus species. In this study, we employ Ecological
Niche Modelling (ENM) and spatiotemporal phylogeographic reconstruction approaches to evaluate a multitude
of scenarios of connectivity across the Neotropical savannas. We identified three divergent lineages, two of which
have been described previously. The lineages were allocated to a lowland Pithecopus clade, although the re­
lationships among these lineages are weakly supported. Both the ENM and the phylogeographic reconstruction
highlight the occurrence of periods of connectivity among the Neotropical savannas over the course of the
Pliocene and Pleistocene epochs. These processes extended from eastern Amazonia to the northern coast of
Brazil. The findings of the present study highlight the presence of hidden diversity within P. hypochondrialis, and
reinforce the need for a comprehensive taxonomic review. These findings also indicate intricate and highly
dynamic patterns of connectivity across the Neotropical savannas that date back to the Pliocene.

1. Introduction isolation of lineages among distinct microhabitats (Guarnizo and Can­


natella, 2013; Rodríguez et al., 2015), and often leads to reduced gene
Amphibians are generally considered to have a limited intrinsic flow and increased lineage differentiation, which further reinforce these
dispersal capacity (Smith and Green, 2005) and tend to have well- patterns. However, philopatry is not a constant in the anurans, and some
defined habitat preferences. Surprisingly, however, some tropical spe­ species of the family Bufonidae may present a tendency for active, long-
cies are relatively widespread (Wynn & Heyer, 2001). The heterogeneity distance dispersal (Fonte et al., 2019), and at least one species of the
of habitats in tropical regions tends to accentuate the specialisation and family Leptodatylidae (Pseudopaludicola falcipes) appears to have high

* Corresponding authors.
E-mail addresses: rafaelfelixm@gmail.com (R.F. Magalhães), elisakramos@gmail.com (E. K. S. Ramos), azebandeira@gmail.com (L.N. Bandeira), johnny.sf@
gmail.com (J.S. Ferreira), fewerneck@gmail.com (F.P. Werneck), marina.anciaes@gmail.com (M. Anciães), danielpachecobruschi@gmail.com (D.P. Bruschi).

https://doi.org/10.1016/j.ympev.2023.107959
Received 12 May 2023; Received in revised form 25 October 2023; Accepted 27 October 2023
Available online 31 October 2023
1055-7903/© 2023 Elsevier Inc. All rights reserved.
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

rates of passive dispersal (Langone et al., 2016). Even so, some recent thigh is wide and covers its full length). Caramaschi (2006) showed that
phylogeographic studies have shown that a number of Neotropical these colour patterns were “congruent” (sic) within each species, which
anuran species previously thought to have extensive geographic ranges suggests a lack of intraspecific variation. Pithecopus hypochondrialis was
may, in fact, be complexes of species, masked by cryptic diversity (e.g., assumed to have an exclusively Amazonia distribution, while the range
Prado et al., 2012; Gehara et al., 2014; Nascimento et al., 2019; Vas­ of the P. azureus morph extends across the Chaco, Pantanal, and Cerrado
concellos et al., 2019; Vacher et al., 2020; Abreu-Jardim et al., 2023). domains.
Pithecopus hypochondrialis is a widespread, cis-Andean leaf frog, Bruschi et al. (2013) revised comprehensively this taxonomic sce­
which occurs in Suriname (the exact type locality is unspecified), Brazil, nario, re-examining the colour patterns and comparing them with DNA
Colombia, Guiana, French Guiana, and Venezuela (Frost, 2023). This sequences and chromosomal data. The results of this phylogenetic
species thrives in a variety of ecological domains, including Amazonia, analysis indicated that P. azureus is restricted to the Chaco and Pantanal
the Cerrado, Guianan (or Roraima) savannas, and the Llanos of regions, and that it has a closer phylogenetic relationship with
Venezuela, where these frogs breed in swamps and ponds (Barrio- P. nordestinus than P. hypochondrialis, a finding supported by other
Amorós, 2009; Bruschi et al., 2013; Guarnizo et al., 2015; Vaz-Silva studies (Faivovich et al., 2010; Haga et al., 2017; Andrade et al., 2020).
et al., 2020). The available evidence indicates that P. hypochondrialis In contrast with the classification of Caramaschi (2006), the populations
occupies open habitats, even within the Amazon rainforest (Barrio- of the Cerrado, together with their Amazonian counterparts, have been
Amorós, 2009), where it is found primarily in savanna enclaves and attributed to P. hypochondrialis rather than P. azureus in Bruschi et al.
forest glades, clearings, and edges (Lescure et al., 1995; Bernarde, 2007; (2013). The authors identified three subclades (referred to here as A, B,
Barrio-Amorós, 2009; Lima et al., 2017). Given the current disjunct and C) within P. hypochondrialis. Subclade A included samples from
distribution of the savannas of South America, which include both Belterra, in the Brazilian state of Pará, and Alta Floresta (Mato Grosso
southern and northern blocks, as well as isolated fragments scattered state, Brazil). Subclade B encompassed samples from Chapada dos
within the Amazonian rainforest (Silva and Bates, 2002; Werneck et al., Guimarães and Santa Teresinha (both in Mato Grosso), while subclade C
2012; Azevedo et al., 2020), P. hypochondrialis has a discontinuous consisted of the remaining samples from the Brazilian states of Amapá,
distribution. This geographic discontinuity may contribute to the po­ Bahia, Maranhão, Minas Gerais, Pará, and Tocantins. The phylogenetic
tential isolation and eventual genetic structuring of its lineages. relationship of these subclades was (A, (B, C)).
Species of many other taxonomic groups have a similarly disjunct In particular, Bruschi et al. (2013) reported a broader variation in
distribution across the Neotropical savannas, including birds (e.g., Silva colouration patterns than that found by Caramaschi (2006), and
and Bates, 2002; Lima-Rezende et al., 2019), lizards (e.g., Avila-Pires, described four distinct phenotypes. These phenotypes presented no
1995; Martins et al., 2021), and trees (e.g., Buzatti et al., 2017; Buzatti systematic geographical structuring, which contradicted the differenti­
et al., 2018). This recurring pattern may be related to the historical ation of P. hypochondrialis and P. azureus by these traits. Bruschi et al.
connectivity among these regions, especially during the Quaternary (2013) also found variation in the distribution of the Nucleolar Orga­
glacial cycles (Silva and Bates, 2002). Three primary corridors of con­ niser Region (NOR) among the P. hypochondrialis populations examined
nectivity between the southern and northern Neotropical savannas have from subclades A, B, and C. In particular, they were able to discriminate
been postulated (Webb, 1991; Silva and Bates, 2002): (i) an Andean subclade A from the other subclades due to the presence of NORs on the
corridor through the cis-Andean slopes; (ii) a coastal Atlantic corridor, short arm of chromosome pair 8, although they interpreted this as
connecting the blocks of savanna along the northern coast of Brazil; and intraspecific variation linked to the population structure of the species.
(iii) a central Amazonian corridor, which assumes that the retraction of The NORs are valuable chromosomal markers, which are pivotal for the
the forest during glacial periods favoured the expansion of savannas diagnosis of many anuran species (Schmid et al., 2014) and the identi­
through central Amazonia, forming a corridor of non-forest habitats that fication of independent evolutionary lineages (Nascimento et al., 2019).
connected previously isolated blocks of savanna. While most of the ev­ More recently, Haga et al. (2017) described the new species
idence favours the Andean and Atlantic coast hypotheses (Silva and P. araguaius, which is closely related to P. hypochondrialis. This species
Bates, 2002; Werneck et al., 2012), recent research has pointed to the corresponds to subclade B in the study of Bruschi et al. (2013). To
potential existence of central Amazonian corridors during the Quater­ support their description, Haga et al. (2017) contrasted the adult
nary (Bueno et al., 2017; Buzatti et al., 2018), although their timing and morphometry, bioacoustic traits, and mitochondrial DNA markers of
configuration are not well understood, in part due to the absence of this new taxon with samples of P. hypochondrialis, classified as
macrofossils and palynological records that would validate the models “P. hypochondrialis North” and “P. hypochondrialis South” from Ama­
(Azevedo et al., 2020). Dowsett et al. (2016) also projected the potential zonia and the Cerrado, respectively. Their findings indicate that
occurrence of savannas in present-day eastern Amazonia, during the P. araguaius can be distinguished from P. hypochondrialis by statistically
Pliocene (mid-Piacenzian, ~ 3 Ma), which represents an additional significant differences in its morphometry and advertisement calls
hypothesis, on an older corridor (Azevedo et al., 2020). The phylo­ (Haga et al., 2017). Given the overlap in the morphometric and acoustic
geography of P. hypochondrialis thus offers an opportunity to decipher attributes of P. araguaius and P. hypochondrialis, however, these closely-
the possible centre of species diversification and test alternative hy­ related species cannot be distinguished reliably based solely on their
potheses with regard to the ancient corridors connecting the Neotropical phenotypic traits. As a result, they are considered to form a complex of
savannas. morphologically and acoustically cryptic species, referred to here as the
Few studies have investigated P. hypochondrialis from a broad P. hypochondrialis species complex. An alternative position would be to
geographical perspective. Caramaschi (2006) examined the external consider them as synonymous taxa.
morphometry and colour patterns of specimens from Argentina, Bolivia, Remarkably, Haga et al. (2017) did not sample any individuals from
Brazil, Colombia, and Suriname, which were all attributed to the nom­ Bruschi et al.’s (2013) subclade A, which raises the possibility of the
inal species P. hypochondrialis at that time. This study identified two existence of at least one more candidate species within the
colour morphotypes, which had previously been treated as two distinct P. hypochondrialis species complex. The phenotypic variation found
species – P. hypochondrialis and P. azureus – and revalidated the latter among these lineages is somewhat confusing. While Bruschi et al. (2013)
taxon. In this analysis, P. hypochondrialis was distinguished from demonstrated that conventional colouration patterns are not diagnostic
P. azureus by the presence of a white stripe on the upper lip, which ex­ of P. hypochondrialis, they did identify differences in the frequency of the
tends to the lower edge of the lower eyelid, and a narrow green stripe on morphotypes among the lineages, possibly along a longitudinal axis (see
only 2/3 to 3/4 of the distal portion of the upper surface of the thigh (in Fig. 1 in Bruschi et al., 2013). By contrast, Haga et al. (2017) reported
P. azureus, the white stripe on the upper lip does not reach the lower morphometric differences along a latitudinal axis, with progressively
edge of the lower eyelid, and the green stripe on the upper surface of the larger individuals moving northwards within the species’ range.

2
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

These contrasting patterns in the distribution of the phenotypes, and Information on the primers used to amplify each fragment is provided in
the lack of an integrative species delimitation (sensu Yeates et al., 2011), Table S3. The Polymerase Chain Reactions (PCRs) were run in a total
raise a number of questions with regard to the taxonomic arrangement volume of 25 µL containing 20–100 ng of the DNA template, 1X PCR
of the species complex and the biogeographic history of its closely- buffer, 1.5 mM MgCl2, 7 pmoles/µl of the forward and reverse primers,
related lineages. In this context, the present study tests alternative hy­ 1 mM of dNTPs, and 1U of Taq polymerase. The temperatures and
potheses, i.e., whether P. hypochondrialis is a single, widespread species annealing times for each fragment are provided in Table S3. The
or a cryptic species complex. For this, we employed coalescent and amplicons were purified with exonuclease I (10 units) and SAP (1 unit)
integrative species delimitation methods to test the evolutionary inde­ to remove the unincorporated dNTPs and primers. The amplified frag­
pendence (sensu Simpson, 1951) of the different lineages. We also ments were sequenced in an ABI PRISM™ 3100 Genetic Analyzer
investigated the extent to which the morphological variation observed (Applied Biosystems, Foster City, CA, USA) in both directions, using the
in the populations reflects the genetic structure found within the original amplification primers and BigDye™ terminator v.3.1
P. hypochondrialis complex, which includes P. azureus in some compar­ sequencing chemistry according to the manufacturer’s protocol
isons, due to the apparent morphological similarities of this species (Applied Biosystems, Foster City, CA, USA). We employed CodonCode
(Bruschi et al., 2013). Aligner 7.1.1 to verify the base calls and to correct the base-calling er­
To address these questions, we investigated the intrinsic genetic rors found in the sequences.
variation and genealogical history of the species complex using multi­ We aligned the fragments with MAFFT v.7.427, which was imple­
locus mitochondrial and nuclear datasets in a comprehensive integrative mented in the online service (Katoh and Standley, 2013; Katoh et al.,
approach. We generated novel sequence data and incorporated se­ 2019). We used the automatic strategy for coding fragments (i.e., ND2,
quences available in public databases to cover the geographic distribu­ RHOD, and SIA-H1), while for the 16S gene, we used the Q-INS-i algo­
tion of the species complex as extensively as possible. We also re- rithm (Katoh and Toh, 2008) set at the default parameters. Finally, we
evaluated the colouration patterns examined by Bruschi et al. (2013) reconstructed the haplotype phases of the nDNA fragments using PHASE
and integrated the morphometric variation of the specimens into our v.2.1.1 (Stephens et al., 2001) for some of the subsequent analyses,
species delimitation analyses. Ultimately, the present study extends the considering a Posterior Probability (PP) threshold of 0.7 to accept the
application of phylogeographic analyses and paleo-distribution model­ estimated phases. The SeqPHASE webtool (Flot, 2010) was used to
ling to a Plio-Pleistocene temporal scale. Through this approach, we aim interconvert the input/output of PHASE.
to decipher the historical connectivity of the savanna corridors of South
America, which may have played a fundamental role in the establish­ 2.2. Morphometric and morphotyping data
ment of the current distribution of the P. hypochondrialis species
complex. We assessed 18 morphometric characters in 151 adult male speci­
mens of P. hypochondrialis and P. araguaius, which were obtained from
2. Material and methods the herpetological collections mentioned above (Table S4). At least one
gene fragment was also sequenced in 79 % of these individuals. We
2.1. DNA sampling and editing measured the Snout-Vent Length (SVL), First Finger Length (FFL), Head
Length (HL), Head Width (HW), Eye Diameter (ED), Tympanum Diam­
We collected 131 tissue samples of P. hypochondrialis and P. araguaius eter (TD), Eye-Nostril Distance (END), Nostril-Snout Distance (NSD),
from 24 localities, together with 10 samples of P. azureus from four sites Inter-Nostril Distance (IND), Inter-Orbital Distance (IOD), and Tibia
(Table S1; see Results). We collected the latter samples due to their Length (TL) following the protocol of Duellman (1970). We also deter­
morphological similarity and geographic proximity to our study species mined the Arm Length (AL), Forearm Length (FAL), Forearm Width
(Faivovich et al., 2010; Bruschi et al., 2013; Haga et al., 2017). The (FAW), Hand Length (HAL), Thigh Length (THL), and Foot Length (FL)
tissue samples were collected from euthanised specimens following the following Heyer et al. (1990), as well as the First Toe Length (FTL),
application of an anaesthetic (5 % Lidocaine) to the skin, in accordance which we defined as the distance between the proximal margin of the
with the recommendations of the Herpetological Animal Care and Use inner metatarsal tubercle and the distal apex of the first toe. We acquired
Committee (HACC) of the American Society of Ichthyologists and Her­ the measurements (in millimetres) using a needlepoint calliper with an
petologists (available at: https://www.asih.org/publications). The accuracy of 0.01 mm.
collection of samples was approved by SISBIO/Instituto Chico Mendes We followed the approach of Lleonart et al. (2000) to eliminate the
de Conservação da Biodiversidade, through collecting permit number potential allometric effects in the data. This involves regressing the
14468-1/14468-4. Voucher specimens or tissue samples were deposited log10-transformed variables against a size proxy determined by a Prin­
in the Célio Fernando Baptista Haddad Amphibian Collection (CFBH) of cipal Components Analysis (PCA). This proxy is the variable with the
the Universidade Estadual Paulista in Rio Claro, São Paulo, Brazil; the highest loading value obtained by the first principal component. The
Prof. Adão José Cardoso Museum of Zoology (ZUEC) at Universidade eigenvectors of the variance–covariance matrix were obtained using the
Estadual de Campinas in São Paulo, Brazil; the Shirlei Maria Recco- ‘prcomp’ function in the vegan v.2.5–6 package of the R software
Pimentel tissue collection (SMRP) at Universidade Estadual de Campi­ (Oksanen et al., 2016), with the variable FAL being selected as the proxy
nas; the Zoological collection of Universidade Federal de Goiás (ZUFG) for the analysis. We then ran pairwise linear regressions between the
in Goiás, Brazil; and the amphibian collection from Museu de Zoologia FAL and the other variables using the stats package in R v.4.0.0 (R Core
da Universidade de São Paulo (MZUSP), in São Paulo, Brazil. We also Team, 2020). All the subsequent analyses were conducted using the
obtained sequences of at least one individual of each Pithecopus species residuals of these regressions. Finally, we calculated the Variance
(except P. gonzagai Andrade et al., 2020) from GenBank (Table S2), as Inflation Factors (VIFs) among the corrected variables, to estimate the
well as one individual of Callimedusa tomopterna, and sequences of multicollinearity of the data, in the car package v.3.0–7 (Fox & Weis­
P. azureus and P. hypochondrialis from locations not represented in our berg, 2018) of the R software. Only traits with VIF values smaller than 5
sampling, including the P. hypochondrialis type locality (Table S3). (Akinwande et al., 2015) were retained for all subsequent analyses but
The genomic DNA was extracted from the samples of liver and PERMANOVA (see section 2.4.3).
muscle tissue (preserved in 95 % ethanol) using the TNES method (see In addition to morphometric traits, colouration patterns have been
Bruschi et al., 2012). We amplified and sequenced fragments of two employed traditionally in taxonomic studies of the subfamily Phyllo­
mitochondrial (mtDNA) genes – NADH dehydrogenase subunit 2 (ND2) medusinae (Caramaschi et al., 2006). While some studies (e.g., Bruschi
and the ribosomal 16S gene (16S) – and two nuclear (nDNA) loci – et al., 2013; Brunes et al., 2014) have raised doubts with regard to the
Rhodopsin exon 1 (RHOD) and Seven in Absentia Homolog 1 (SIA-H1). efficacy of these features for the diagnosis of taxa, they can provide

3
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

additional insights into the differentiation of species, when an appro­ sequences were missing in the ND2 dataset (vs. 17 % in the case of 16S;
priate statistical approach is adopted. Given this, we re-evaluated the Tables S1-2). We identified the unique haplotypes in the haplotypes
four colour morphs defined by Bruschi et al. (2013) and assigned each of v.1.1.2 package of the R software (Aktas, 2020) by using the haplotype()
the 179 specimens of P. hypochondrialis, P. araguaius, and P. azureus function with the indels=“sic” option, and then removed the equal
analysed here (Table S4) to one of these morphotypes. Pithecopus azureus haplotypes before a second run with the indels=“missing” option. This
was included in these comparisons specifically to assess the extent of the step is important because the accumulation of short branches can lead to
differentiation between this species and P. hypochondrialis, considering an overestimate of candidate species by the GMYC algorithms (see
their complex taxonomic history. We classified the colour morphs section 2.4.1) (Reid and Carstens, 2012; Talavera et al., 2013).
following the exact procedure outlined by Bruschi et al. (2013), as We generated a set of ultrametric gene trees from these filtered se­
follows: quences using the BEAST 2 software (Bouckaert et al., 2019). For this
Morphotype 1: Narrow white stripe on the upper lip extending to the analysis, the substitution model was identified using bModelTest, as
lower eyelid together with the presence of a discontinuous, wide green above, with the Yule tree and log-normal clock models. This combina­
stripe along 2/3 to 3/4 of the length of the upper surface of the thighs. tion of priors resulted in a high percentage of correct hits in the GMYC
Morphotype 2: White stripe on the upper lip extending to the lower performance test, based on empirical data from butterflies (Talavera
eyelid together with the presence of a wide green stripe along the full et al., 2013). The analysis was run twice, with 1 × 107 generations in
length of the upper surface of the thighs. each case, and thinning to 1 × 103. The convergence, mixing, and
Morphotype 3: White stripe on the upper lip extending to the lower minimum ESS were all verified as above. The results of these runs were
eyelid together with a green stripe absent on the upper surface of the combined and MCC tree was annotated after a 5 % burn-in.
thighs.
Morphotype 4: Thin white stripe on the upper lip that does not 2.3.3. Nuclear haplotype networks
extend as far as the lower eyelid. Wide green stripe along the full length We constructed nuclear haplotype networks for comparative pur­
of the upper surface of the thighs. poses. For this, we generated Maximum Likelihood (ML) gene trees for
the RHOD and SIA-H1 fragments using MEGA X v.10.2.6 (Kumar et al.,
2.3. Phylogenetic analyses 2018) under the GTR model. The initial trees for the heuristic search
were obtained automatically by applying the Neighbor-Joining and
2.3.1. Concatenated phylogeny BioNJ algorithms to a matrix of pairwise distances estimated using the
Given the difficulty of distinguishing P. araguaius, P. azureus, and Maximum Composite Likelihood approach, with the topology with the
P. hypochondrialis based on their external morphology (Bruschi et al., highest log-likelihood value being selected. The ML gene trees were
2013; Haga et al., 2017), we cannot rule out altogether the possibility converted into haplotype genealogies using the HaploViewer program
that some of the specimens obtained from the biological collections have (Salzburger et al., 2011).
been misidentified. Given this, we constructed a tree concatenating all
the gene fragments of each species into a super-gene alignment before 2.3.4. Species tree dating
running the lineage delimitation analyses, in order to verify the taxo­ To estimate the relationships and divergence times among the spe­
nomic identity of each individual. To do this, we included specimens cies identified in the P. hypochondrialis species complex (see Results), we
from the type localities of P. araguaius (Pontal do Araguaia, Mato Grosso constructed a species tree in the StarBEAST2 v.0.15.5 package (Ogilvie
state, Brazil) and P. hypochondrialis (Suriname), as well as specimens et al., 2017) of the BEAST2 software. For this, we linked the 16S and
that can be attributed unequivocally to P. azureus (from localities in ND2 fragments in the same mtDNA gene tree but estimated their sub­
Argentina and Brazil, within the basins of the Paraná and Paraguay stitution models and clocks independently. The trees and parameters of
rivers; Caramaschi, 2006), as taxonomic references. the nDNA fragments were estimated independently. The substitution
To compile the phylogeny, we first excluded all the individuals that model of each fragment was co-estimated using the bModelTest package
lacked sequences for both mitochondrial fragments. We also included (Bouckaert and Drummond, 2017), with the set of test models being
representatives of all the remaining Pithecopus species except defined as ‘allreversible’. The tree prior was set as the birth-and-death
P. gonzagai, and rooted the tree at Callimedusa tomopterna, given that process (Gernhard, 2008), and the species tree was rooted at
Callimedusa is the sister genus of Pithecopus (Faivovich et al., 2010; P. azureus. Given the lack of phylomedusine fossils with which to cali­
Duellman et al., 2016). We ran the Bayesian inference in BEAST v. 2.6.2 brate the species tree, we used the substitution rate of the ND2 gene
(Bouckaert et al., 2019). We concatenated all the unphased fragments (i. (0.957 % per lineage per million years; Crawford, 2003), which was
e., linking all the trees) while estimating the site and clock models of used as the reference to estimate the rates of the other fragments, with
each one independently, to increase the accuracy of the reconstruction an uncorrelated relaxed log-normal clock being assumed for all the
of the branch lengths. The site model was co-estimated with the fragments (Drummond et al., 2006). We ran the analysis twice, with 1 ×
phylogenetic analyses using the bModelTest package v. 1.2.1 (Bouckaert 108 generations each, thinning to 1 × 104, and burnin 1 × 106.
and Drummond, 2017), with the set of test models being defined as Convergence between runs, mixing of the parameters, and ESS were
‘allreversible’. The strict clock was set for all four fragments, with the verified in Tracer. The results of both runs were merged and the MCC
clock rate being fixed at 1.0 for the 16S gene, with all the others being tree was annotated using the LogCombiner and TreeAnnotator.
estimated relative to this gene. Finally, the tree prior was set to the Yule
Model. We ran the analysis twice, with 5 × 107 generations each, 2.4. Species delimitation
thinning to 5 × 103. Convergence between the runs, the mixing of the
parameters, and the minimum Effective Sample Sizes (ESS > 200) were 2.4.1. Molecular species discovery and validation
verified in Tracer v. 1.7.1 (Rambaut et al., 2018). The results of both We delimited the lineages within P. hypochondrialis by applying a
runs were combined, with the first 10 % of the trees being discarded as two-step iterative procedure, consisting of discovery and validation
burn-in, and the Maximum Clade Credibility (MCC) tree being anno­ approaches (Carstens et al., 2013), to the molecular data. In the dis­
tated using the LogCombiner and TreeAnnotator post-processing tools, covery step, we estimated the number of putative taxonomic units using
respectively (Bouckaert et al., 2019). the General Mixed Yule-Coalescent (GMYC) model (Pons et al., 2006).
The GMYC analysis was implemented using two distinct algorithms. The
2.3.2. Mitochondrial gene trees Maximum Likelihood version (ML-GMYC) was applied to the MCC gene
Only the 16S sequences were used to generate the ultrametric tree (see section 2.3.2), using the single-threshold method implemented
mtDNA gene trees, given that a large proportion (38.3 %) of the in the splits package v.1.0–19 of the R software (Fujisawa and

4
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Barraclough, 2013; R Core Team, 2020). In addition, to account for the assumptions about the taxonomic identity of the individuals. We ran the
uncertainties in the topology and branch lengths estimation, we applied first analysis with Kmin = 1 and Kmax = 12 (i.e., the optimal number of
the Bayesian version of the model (bGMYC) to a sample of 2.5 × 103 genetic clusters identified in the first iteration of the GMYC: see Results).
random trees generated by BEAST 2 (see section 2.3.2). This analysis We conducted 10 independent runs, each with 1 × 107 generations, 1 ×
was conducted in the bGMYC v.1.0.2 package of the R software (Reid 103 thinning, and a burn-in period of 150 first results before thinning.
and Carstens, 2012; R Core Team, 2020), using the single-threshold We checked for mixing and convergence by inspecting the posterior
method and limiting the boundaries for the number of putative species density of each run and comparing the results across the runs. Once we
to a threshold between one and the maximum value of the credibility had determined the best K value, we conducted an additional run, for
interval of the entities delimited by the ML-GMYC analysis. which we kept the K fixed at its optimal estimate. We used this step to
Talavera et al. (2013) reported an improvement in precision when generate more precise morphometric membership probabilities for the
the taxonomic operational units are represented by only a single or a few individual samples (Guillot et al., 2005). We then generated the Q-plot
haplotypes. Given this, we ran the discovery step twice. In the first step, for this final run using DISTRUCT version 1.1 (Rosenberg, 2004). To
we constructed the mtDNA gene tree using all the unique haplotypes. In determine whether the morphometric data could discriminate inde­
the second iteration, we randomly selected one of the most complete pendently the lineages validated by the genetic species delimitation
sequences delimited per lineage among the sequences with the greatest analysis (see above), we applied a Permutational Analysis of Variance
length (Table S5), estimated new gene trees, and ran the ML-GMYC and (PERMANOVA) to the morphometric traits in Euclidian distance, using
bGMYC analyses once again. The results of this second iteration were the vegan v. 2.5–7 package of the R software (Oksanen et al., 2016).
used in the validation step. To compare the results of the GMYC itera­ Finally, to test for patterns of variation in the frequency of the
tions, we estimated the uncorrected p-distances within and between the different colour morphotypes among the candidate species, we applied
16S lineages delimited by each step. These distances were generated Pearson’s χ2 test in R stats v. 3.5.0. We included the P. azureus specimens
using MEGA X, with all the ambiguous positions (i.e., missing data and in this analysis to test Caramaschi’s (2006) hypothesis that individuals
gaps) being removed from each sequence pair (pairwise deletion of this species can be differentiated from P. hypochondrialis based on
option). their colouration patterns (i.e., colour morphotypes 4 and 1,
The validity of the lineages discovered during this analysis was tested respectively).
using the mtDNA and nDNA datasets. For this, we used the STACEY
v.1.2.5 package (Jones, 2017) of the BEAST 2 software, grouping in­ 2.4.4. Integrative species validation
dividuals that were discovered in the previous step as putative species. We employed an integrative validation analysis, based on the inte­
Each locus was now set to have its own independent substitution model, grated Bayesian Phylogenetics and Phylogeography model, which we
estimated by bModelTest. The clock rates were set as strict and inde­ implemented in iBPP v.2.1.3 (Solís-Lemus et al., 2015). We used both
pendent for each fragment. The 16S and ND2 fragments were linked in the molecular and the morphometric data, following the VIF correction,
the same mtDNA gene tree. The inheritance scalars were set at 0.5 for to satisfy the assumptions of the analysis. As the species tree retrieved
the mtDNA, and at 2 for the nDNA genes. We set the collapse height of low support for the relationships among the lineages (see Results), we
the priors as 1 × 10-3, their collapse weight as ~ uniform (0.1), the ran distinct analyses using all the three possible species trees, and tested
population prior scale as ~ lognormal (M = -7, S = 2), the growth rate as the divergence scenario of the putative species recovered by the second
~ lognormal (M = 5, S = 2), and the relative death rate as ~ β (α = 1, β iteration of the GMYC. Four distinct gamma prior combinations were
= 8). The analysis was run twice, with 1 × 108 generations in each case, used for the theta (θ) and tau (τ) parameters, with two replicates each,
and thinning to 1 × 104. The convergence, mixing, and minimum ESS including (i) large ancestral populations and deep divergence (θ ~ G (2,
were all verified as above. The results of these runs were combined and 100), τ ~ G (2, 200)), (ii) small ancestral populations and shallow
MCC tree was annotated after a 10 % burn-in. The best number of divergence (θ ~ G (2, 1000), τ ~ G (2, 2000)), (iii) small ancestral
candidate species was determined using the speciesDA program (Jones, populations and deep divergence (θ ~ G (2, 1000), τ ~ G (2, 200)), and
2017) with a burn-in of 10 %, and collapse heights of 1 × 10-4, 5 × 10-4, (iv) large ancestral populations and shallow divergence (θ ~ G (2, 100),
7.5 × 10-4, and 1 × 10-3. This set of heights was adopted to verify the τ ~ G (2, 2000)). The priors for the continuous morphometric trait data
robustness of the delimitation. were set as in Solís-Lemus et al. (2015). We used the Dirichlet prior
(Yang and Rannala, 2010) for the locus-rate parameters. We chose an
2.4.2. Isolation-by-distance test automatic adjustment in the MCMC step length, and ran all the analyses
As Isolation By Distance (IBD) may lead to an overestimate of the with 2 × 105 generations and a burn-in of 2 × 104 generations.
number of candidate species in molecular delimitation methods
(Chambers and Hillis, 2020; Mason et al., 2020), we applied the Mantel 2.4.5. Heuristic species delimitation
test (Mantel, 1967) to the pairwise FST and geographical distance To assess the strength of the divergence among the lineages validated
matrices of the 16S sequence data. This analysis was implemented in the by the iBPP, we conducted a heuristic analysis using the genetic diver­
Arlequin v.3.5.2.2 software (Excoffier and Lischer, 2010). We selected a gence index (gdi) of Jackson et al. (2017). We initially ran the A00
specific subset of the 16S sequences for the IBD analysis (Table S1) as a analysis in the iBPP framework to obtain the posterior distributions for
strategy to reduce the impact of incomplete sequences on our analysis. the τ and θ parameters across the three different tree topologies with the
We estimated the linear distances between sampling sites using the prior combinations of θ and τ used in the iBPP. For this, we applied the
Geographic Distance Matrix Generator v.1.2.3 (GEODIS) platform of the same MCMC settings as those used in the integrative species validation.
American Museum of Natural History (Posada et al., 2000), using the We subsequently used these distributions to calculate the gdi for all the
sample coordinates. The runs consisted of 1 × 104 permutations and potential species divergence events based on our guide trees. We
were applied to the sequences representing the entire distribution of the calculated the gdi values using the R function available at https://github.
species complex and separately for each lineage discovered during the com/melisaolave/mafalda. In this approach, the study populations are
analysis. classified as distinct species if their gdi values exceed 0.7, while gdi
values below 0.2 indicate that they should be considered to belong to the
2.4.3. Morphological analyses same species. Intermediate gdi values, ranging from 0.2 to 0.7, indicate
We used morphometric data to estimate separately the optimal an ambiguous species status (Jackson et al., 2017).
number of clusters using the phenotypic model implemented in the
Geneland v.4.0.6 package of the R (Guillot et al., 2012; R Core Team,
2020), without incorporating the spatial model or making any a priori

5
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

2.5. Paleodistribution models and climatic stability across geological M2 (ca. 3.3 Million years ago – Ma), (ii) the mid-Pliocene warm period
periods (3.205 Ma), (iii) MIS19 (ca. 787 thousand years ago – ka), (iv) the Last
Interglacial – LIG (ca. 130 ka), (v) the Last Glacial Maximum – LGM (ca.
We modelled the Ecological Niche (ENM) of the species complex to 21 ka), (vi) Heinrich Stadial 1 (11–14.7 ka), (vii) the early Holocene
estimate both its current potential distribution (i.e., the Species Distri­ (11.7–8.326 ka), (viii) the mid Holocene (8.326–4.2 ka), and (ix) the late
bution Model, SDM) and its historical distribution during nine periods in Holocene (4.2–0.3 ka).
the past (0.3 ka–3.3 Ma, see below), as well as mapping the areas of The ENMs were constructed using the Maximum Entropy algorithm,
suitable climate available over time (areas of stability). MaxEnt (Phillips et al., 2006), which was implemented in the KUENM
To this end, the totality of the occurrence records of package v.1.1.7 (Cobos et al., 2019) in R v.4.0.0. The MaxEnt algorithm
P. hypochondrialis species complex was used to construct the niche generates predictions from presence-only data, based on the principle of
models using the subsequent modelling routines. The occurrence re­ maximum entropy, which assumes that the best approximation for an
cords were compiled from the herpetological collections consulted unknown distribution is the one that satisfies any constraint on its dis­
during the study Bandeira et al. (2021) and are listed in Table S6. We tribution (Phillips et al., 2006; Elith et al., 2011). This algorithm is
also verified the potential existence of complementary records in the considered to be extremely efficient and is widely used to model species
relevant published papers to include data points that could be assigned distributions.
unambiguously to a precise locality via the GPS coordinates provided in The final models were generated using 10 independent replicates,
the publication. After obtaining, verifying, and validating each occur­ while 50 % of the points were used for testing (Fig. S1), using the
rence record, our final set of unique geo-referenced localities included Bootstrap method (Pearson, 2010). We chose the logistic output to
296 records of P. hypochondrialis species complex, which represents a represent the ENMs in geographic space (SDM), with the suitability of
satisfactory coverage its known geographic range (Table S6). each pixel (geographic cell) ranging from 0 (totally inadequate condi­
To compile the current and past climate patterns, we assembled 17 tions) to 1, i.e., maximum suitability (Phillips et al., 2006). The per­
layers of climatic variables with continuous data on temperature and formance of the candidate models was evaluated hierarchically using
precipitation that are closely related to the ecological and physiological three independent, but complementary criteria: significance, predictive
tolerances of anurans (Duellman & Trueb, 1994; Wells, 2007). All the ability, and complexity. The models were then filtered to detect the
variables were obtained from the Paleoclim online bioclimatic dataset statistically significant ones, and the “low-omission criterion” (the Area
(https://paleoclim.org/, Otto-Bliesner et al., 2006; Dolan et al., 2015; Under the Curve [AUC] method; Phillips et al., 2006) was applied to
Hill, 2015; Fordham et al., 2017; Brown et al., 2018) at a resolution of further reduce the set of models. Finally, the significant models with the
2.5′ (~5 km). For the present day, bioclimatic parameters were derived lowest omission rate whose delta AICc values were lower than two were
from the monthly min, max, mean temperature, and mean precipitation selected as the best models (Cobos et al., 2019).
values, to determine annual trends (e.g., mean annual temperature, To project the SDM of the P. hypochondrialis species complex into the
annual precipitation), seasonality (e.g., the annual range in temperature past, we converted the suitability maps (with values ranging from 0 to 1)
and precipitation), and extreme or limiting environmental factors (e.g., to binary presence-absence maps using a “cut-off” threshold value. As
temperature of the coldest and warmest months, and the precipitation of recommended by Liu et al. (2013), we converted the SDMs using the
the rainiest and driest quarters), for the 1967–2013 time-series. For past “Maximum Test Sensitivity plus Specificity – MTS + S” criterion as the
periods, we decided to assemble paleo-climatic simulations representing threshold value. This minimizes the areas identified erroneously as un­
nine periods ranging from the Pliocene to the Holocene: (i) the Pliocene, suitable (false negatives) in the candidate models, as well as those

Fig. 1. Maximum clade credibility tree obtained from the Bayesian phylogenetic inference of the concatenated mtDNA (16S, ND2) and nDNA (RHOD, SIA-H1)
fragments, showing the relationships among the Pithecopus species. The information contained within the parentheses includes the acronym and the registration
number of the collection in which the specimen is deposited, a two-letter ISO (International Organisation for Standardisation) 3166 standard country code, the state
or department, and the municipality of registration. The circles at the branch nodes show their respective posterior probabilities (PP), as indicated in the legend.
Nodes with PP > 0.95 (strong support) are denoted by black circles. Please see Fig. 2 for further insights into the relationships within the P. hypochondrialis
species complex.

6
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

(caption on next page)

7
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Fig. 2. Maximum clade credibility tree obtained from the Bayesian phylogenetic inference of concatenated mtDNA (16S, ND2) and nDNA (RHOD, SIA-H1) frag­
ments, showing the relationships among the specimens in the Pithecopus hypochondrialis species complex. The information contained within the parentheses includes
the acronym and the registration number of the collection in which the specimen is deposited, a two-letter ISO (International Organisation for Standardisation) 3166
standard country code, the state or department, and the municipality of registration. Branches with highlited information in bold represent topotypic individuals of
P. araguaius and P. hypochondrialis. The circles at the branch nodes show their respective posterior probabilities (PP), as indicated in the legend. Nodes with PP > 0.95
(strong support) are denoted by black circles. Nodes with PP < 5 % are not highlighted with circles. The numbers within the white vertical bars with black borders on
the right side of the tree refer to the lineages identified during the initial iteration of the general mixed Yule-coalescent (GMYC) analysis (see Results; Fig. S3). The
colour-coded vertical bars on the right indicate the lineages validated by the molecular delimitation analyses (see Results). The geographical distribution of these
lineages is depicted using colour-coded minimal convex polygons in the map in the lower left corner. The circles outlined in black on this map represent the type
localities of P. araguaius (orange) and P. hypochondrialis (blue). For the relationships within the outgroups, see Fig. 1. (For interpretation of the references to colour in
this figure legend, the reader is referred to the web version of this article.)

lacking predicted suitable habitat (false positives), which has a higher 3. Results
risk of overestimating distributions rather than missing important
habitats. 3.1. Concatenated phylogenetic analysis and haplotype networks
Finally, each of the SDMs was intersected to obtain a final model
representing the distribution of climatically stable areas for the occur­ The concatenated tree has two main Pithecopus clades (Fig. 1). One
rence of the species complex over time. In geographical terms, this clade includes P. palliatus, P. rohdei, P. megacephalus, P. rusticus,
model represents all the sites or regions that potentially remained P. oreades, P. centralis, and P. ayeaye. While the inclusion of P. palliatus in
occupied by P. hypochondrialis throughout all the periods analysed, i.e., this clade is poorly supported (PP = 0.25; Fig. 1), the remaining species
that remained suitable for the species and can be identified as a climatic form a highly supported (PP = 1) monophyletic group. The other clade is
refugium (e.g., Carnaval and Moritz, 2008; Werneck et al., 2012). strongly supported (PP = 0.99), and includes P. nordestinus, P. azureus,
P. araguaius, and P. hypochondrialis, with P. araguaius + P. hypochon­
2.6. Spatiotemporal phylogeographic dynamics drialis (or the P. hypochondrialis species complex) as the sister of a clade
composed by P. azureus + P. nordestinus (Fig. 1).
We used a spatiotemporal species tree diffusion approach (Nylinder The monophyly of P. azureus is confirmed and well supported
et al., 2014) to determine which Neotropical savanna corridor hypoth­ (Fig. 1), including individuals from the Pantanal, and the Chaco and
esis best fits our data. This approach is based on the log-normal Relaxed Beni savannas. Three major clades with strong nodal support (PP > 97
Random Walk (RRW) algorithm (Lemey et al., 2010), which was %) can be observed within the P. hypochondrialis complex (Fig. 2). The
implemented in BEAST v.1.8.4 (Drummond et al., 2012). The species largest of these clades includes a P. hypochondrialis sample from the type
trees used as the input were produced by StarBEAST2, as described locality (i.e., Suriname) and specimens collected from north-eastern
above. However, we initially estimated the relationships among the 12 Amazonia and the core Cerrado domains, which we identify as
lineages discovered in the first iteration of the GMYC rather than the P. hypochondrialis here (lineage 3; Fig. 2). This clade is the sister group of
three validated species (see Results, Fig. 2). In this step, we merged the the P. araguaius (lineage 2) + Pithecopus sp. (lineage 1) clades (Fig. 2).
individuals from Santarém with those of lineage 1.2 (see Results). We The P. araguaius clade (lineage 2; Fig. 2) includes a subclade composed
used the ND2 substitution rate (see section 2.3.4) to calibrate the species of topotypes of P. araguaius (Haga et al., 2019) from Pontal do Araguaia
trees, and the substitution rates estimated in the first StarBEAST2 (Mato Grosso) plus samples from Chapada dos Guimarães, in Mato
analysis for the other genes, with a normal distribution (mean = 4.82 × Grosso (PP = 1), and Colombia, whereas the sample from Santa Ter­
10-3, sigma = 5.70 × 10-4 for 16S; mean = 3.06 × 10-4, sigma = 9.20 × ezinha (Mato Grosso) groups with those from Marabá (Pará) in a
10-5 for RHOD; mean = 5.84 × 10-4, sigma = 1.55 × 10-4 for SIA-H1), strongly-supported subclade (PP = 0.99; Fig. 2). It is important to note
setting the clocks as uncorrelated relaxed log-normal (Drummond et al., here that the samples from Colombia and Marabá were identified pre­
2006). Each analysis was run twice, with 1 × 108 generations, 1 × 104 viously as P. hypochondrialis (Fig. 2), and that our findings support the
thinning, and 10 % burn-in. The ESS and convergence between runs revision of the species allocation of these specimens. Finally, the samples
were checked in Tracer, as above. The log and tree files of each gene tree from the northwestern border between Cerrado and Amazon (Paranaíta,
were combined using LogCombiner v.1.8.4, and the MCC gene trees Apiacás, Nova Bandeirantes, and Juína – Mato Grosso) plus Santarém
were computed with TreeAnnotator v.1.8.4 (Drummond et al., 2012). (Pará), which were also identified previously as P. hypochondrialis, were
When three or more geographic records were available for an recovered together in the same clade (lineage 1; Fig. 2).
inferred lineage (i.e., lineages 1.3, 3.1, 3.4, 3.5, and 3.6; see Results), the We obtained sequences of RHOD for 63.8 % of the individuals and
geographic range was represented as a minimum convex polygon, with a SIA-HI for 44.7 % (Tables S1–2). Although we did not find any exclusive
0.5 decimal degree buffer. In the case of the lineages with only one re­ haplogroups for the RHOD fragment, some structuring in the nDNA of
cord (i.e., lineages 1.1 and 3.2), the range was drawn as a circle with a these lineages is indicated by the exclusive haplotypes identified in the
radius of 0.5 decimal degrees around the point. Finally, the geographic SIA-H1 (Fig. S2).
ranges of the lineages with two records (i.e., lineages 1.2, 2.1, 2.2, 2.3,
and 3.3) were represented as an ellipse formed by a 0.5 decimal degree
3.2. Molecular species delimitation
buffer around a line connecting the two points. We used the tutorial
provided by Nylinder et al. (2014) to configure the input analysis from
The first iteration of the ML-GMYC analysis delimits 12 distinct en­
the posterior sample generated by the StarBEAST2 analysis (see section
tities (Fig. S3a), with a confidence interval of 9–15. The null hypothesis,
2.3.4) and the geographic distribution of the lineages. We ran the
that no distinct species exist in the gene tree, can be rejected, based on
analysis twice, with different starting points for each lineage. Each
the Likelihood Ratio (LR) test = 9.09, p = 0.01. All these lineages are
replicate was run over 5 × 107 generations, with 5 × 104 thinning, and
well supported by the bGMYC, although there is some uncertainty on the
10 % burn-in. We combined the results of the two runs in LogCombiner
differentiation of lineages 3.3 and 3.4 as independent candidate lineages
and annotated the MCC tree using TreeAnnotator. Finally, we used the
(Fig. S3a). The second iteration of the ML-GMYC delimits two entities
MCC tree to compile the spatiotemporal dynamics using Spread v.1.0.7
(Fig. S3b), with a confidence interval of 1–11. In this case, however, the
(Bielejec et al., 2011), and summarised the variation in diffusion rates
null hypothesis of no distinct species in the gene tree cannot be rejected
over time using the TimeSlicer tool (Lemey et al., 2010).
(LR test = 2.41, p = 0.30). On the other hand, the scenarios of one and
two candidate species are both rejected by the bGMYC analysis, in

8
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

favour of three well-supported candidate species (Fig. S3b). Table 1


While the “12 candidate species” scenario results in lineages with Results of Permutational Analysis of Variance (PERMANOVA) showing the
low pairwise genetic distances (range: 0.9–5.6 %) and some overlap variation in the morphometric data from putative species validated in the pre­
between intra- and inter-lineage distances (Appendix A), the “3 candi­ sent study (=groups).
date species” scenario results in lineages with higher pairwise genetic d.f. Sum of Squares Mean Squares F p (>F)
distances (range: 4.2–4.5 %), and a clear gap between the intra- and Groups 2 0.007601 0.003801 1.7029 0.1876
inter-lineage distances (Appendix A). Given this, we proceeded with the Residuals 96 0.21426 0.002232
“3 candidate species” scenario for validation. These three candidate
species are validated by STACEY, with a probability of more than 99 %,
regardless of the collapse height used. We assign the three lineages as The variation in the frequencies of the colour morphotypes among
follows: Lineage 1: Pithecopus sp., Lineage 2: P. araguaius, and Lineage 3: the lineages in the P. hypochondrialis species complex and P. azureus
P. hypochondrialis (Fig. 2). cannot be explained by chance (χ2 = 189.19; d.f. = 9; p < 0.001). All the
We found isolation by distance (IBD) in the 16S dataset, with a individuals of the species P. azureus (n = 17) are morphotype 4. In the
moderate correlation between the pairwise FST values and the case of P. araguaius, morphotypes 1, 2, and 4 represent 10.0 %, 75.0 %,
geographic distances between the sampling points (R2 = 0.3781, p < and 15.0 % of the observed variation, respectively (n = 20). Pithecopus
0.001). We also found moderate IBDs in lineages 1 (R2 = 0.5150, p < hypochondrialis is the species with the greatest diversity of colouration
0.05) and 3 (R2 = 0.3065, p < 0.01), and a strong IBD in lineage 2 (R2 = patterns, given the presence of individuals of all four morphotypes.
0.8329, P < 0.05). Morphotype 1 represents 17.4 % of the individuals in the species,
morphotype 2, 39.4 %, morphotype 3, 4.6 %, and morphotype 4, 38.5 %
3.3. Morphological delimitation (n = 109). All but one of the individuals of Pithecopus sp. presents
morphotype 3, with the other one being morphotype 1 (n = 33). Overall,
The independent variables (i.e., those with VIF < 5) used in the then, the colouration patterns are consistent with the species delimita­
Geneland analysis were FAL, SVL, IOD, END, AL, THL, TL, FL, HL, HW tion outlined above, although they cannot be used as diagnostic
and FTL. This analysis identified three morphometric clusters, with PP characters.
= 0.99. However, these clusters did not coincide with the candidate
species validated by STACEY (Fig. 3). The results of the PERMANOVA 3.4. Dated species tree
further confirm this, showing no significant morphometric divergence
among the three lineages validated in the previous steps (Table 1). The StarBEAST2 analysis recovered Pithecopus sp. as a sister of
Despite the extensive mixing of the morphometric clusters within each P. araguaius plus the P. hypochondrialis clade (Fig. 4). However, the re­
delimited lineage, the morphometric traits appear to be conserved at lationships among these three lineages are weakly supported, as they are
many of the localities analysed (Fig. 3). in the concatenated tree (Fig. 2). The divergence time estimates indicate

Fig. 3. The Q-plot of the morphometric variation (left) of the specimens analysed in the present study for K = 3. The individuals are represented by the horizontal
bars and the colours (i.e., light grey, grey, and dark grey) represent the proportion of the individuals assigned to each cluster. The coloured vertical bars to the left of
the Q-plot represent the individuals that belong to the three candidate species validated by the species delimitation methods: Pithecopus sp. (red), P. araguaius
(orange), and P. hyponchondrialis (blue). The vertical lines or arrows with numbers to the right of the Q-plot indicate the collecting localities of the different in­
dividuals, which are shown on the map on the right, all within Brazil. The Brazilian states area as follow: AP = Amapá; BA = Bahia; GO = Goiás; MA = Maranhão;
MG = Minas Gerais; MT = Mato Grosso; PA = Pará; TO = Tocantins. (For interpretation of the references to colour in this figure legend, the reader is referred to the
web version of this article.)

9
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Fig. 4. Time-calibrated species tree produced by the StarBeast2 analysis. The values above the branches (in black) indicate the mean age of the node, while the blue
bars indicate the 95% credibility intervals of the divergence times. The values under the branch (in red) indicate the posterior probability of each node. Photograph of
a specimen of Pithecopus hypochondrialis from Oxiximiná (in Pará state, Brazil) by Samuel C. Gomides. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

that the most recent common ancestor of the P. hypochondrialis species Pithecopus sp. and P. araguaius in the guide tree where they are sister
complex dates back to the Messinian stage of the Miocene (6.63 Ma, 95 taxa, with any other combination of θ and τ priors. There is no support
%-HPD 8.78–3.60 Ma). Pithecopus araguaius and P. hypochondrialis for any of the three species when deep divergences priors are applied, in
appear to have diverged at the end of the Zanclean stage of the Pliocene any of the other guide trees (Table S8).
(3.79 Ma, 95 %-HPD 5.18–0.90 Ma).

3.6. Paleodistribution models and climatic stability over geological time


3.5. Integrative and heuristic validation
The SDM of the P. hypochondrialis species complex has a high AUC
We ran two replicates of the iBPP for each of three possible phy­ value (0.805), with three bioclimatic variables contributing most to the
logenies: (i) ((Pithecopus sp., P. araguaius), P. hypochondrialis), (ii) model – the seasonality of temperature (bio4, contribution = 32.4 %),
((Pithecopus sp., P. hypochrodrialis), P. araguaius) and (iii) ((P. araguaius, precipitation (bio15, contribution = 15.3 %), and the mean temperature
P. hypochondrialis), Pithecopus sp.). We tested four different scenarios for of the driest quarter (bio9, contribution = 11.1 %). In general, the po­
each possible phylogeny, with four combinations of the parameters of tential distribution of the species complex was ampler in the past than
the species divergence times (τ) and population size (θ), which resulted that predicted for the present day (Fig. 5a–j). Under present-day climate
in a set of two replicates of 12 different analyses (Table S7). The three conditions, the predicted potential distribution of P. hypochondrialis
putative species were validated in 20 of these 24 replicates, with PP > encompasses most of the transition zone between the Amazon Forest and
70 % (Table S7). Only four scenarios did not recover all the putative the Cerrado, with relatively high suitability (>0.6), in particular in the
species with PP > 70 %, and these were all in one of the two replicates. northern Cerrado, in the Brazilian states of Tocantins and Maranhão,
Three of these scenarios (Table S7) refer to the test of the phylogeny in and further north within the Amazon biome, including the Brazilian
which P. hypochondrialis (lineage 3) is the sister species of the clade state of Roraima and the open formations – the Llanos – of northern
containing Pithecopus sp. (lineage 1) and P. araguaius (lineage 2), with (i) Venezuela, and central and northwestern Colombia (Fig. 5J).
small ancestral population size and shallow divergence (replicate I.2a), Projections of the current SDMs onto the paleoclimatic conditions of
(ii) large ancestral population sizes and deep divergence (replicate I.4a) Pliocene M2 (~3.3 Ma) indicate the largest and most continuous po­
and (iii) large ancestral population size and deep divergence (replicate tential area suitable for the P. hypochondrialis species complex, with high
I.1b) (Table S7). The fourth scenario that did not recover all the three levels of suitability within the current distribution of the Llanos, central-
species refers to the phylogeny in which P. araguaius is the sister species northern Cerrado and eastern Amazonia (Fig. 5a). This supports the
of a Pithecopus sp. + P. hypochondrialis clade, with a small ancestral conditions for an eastern Amazonian corridor during the Pliocene,
population size and shallow divergence (replicate II.2b). which would have persisted into the warm period of the mid-Pliocene
The results of the gdi vary significantly depending on the combina­ (~3.3 Ma), except for the fragmentation of suitable areas in the cen­
tion of guide trees, θ, and τ priors. In general, the evolutionary inde­ tral Cerrado (Fig. 5b). The models predicted gradual contractions over
pendence of the validated species is uncertain (0.2 < gdi < 0.7; time in the climatically suitable areas towards the transition zone be­
Table S7). The divergence of P. hypochondrialis and P. araguaius, and of tween the Amazon and Cerrado domains (Fig. 5e–j). The climate of the
Pithecopus sp. and P. hypochondrialis are well supported (gdi > 0.7) only northern coast of South America was suitable for the species during
when each of these pairs are sister taxa in the guides trees and with many periods in the past (Fig. 5c, f, h, i), which supports the hypothesis
shallow divergence priors (Table S8). In the case of the genealogical of a coastal Atlantic corridor during the Pleistocene.
divergence between Pithecopus sp. and P. araguaius, there is only strong The map of climatic stability (Fig. 5k) reveals the existence of three
support in the guide tree when they are sister taxa with the combination large climatically stable refugia for P. hypochondrialis over time, corre­
of the priors for shallow divergence and small ancestral populations sponding to the Cerrado/eastern Amazonia in the southeast, the
(Table S8). However, there is no support (gdi < 0.2) for the divergence of Guianan (or Roraima) savannas in the centre, and the Llanos in the

10
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Fig. 5. (a)-(j) Distribution models for the Pithecopus hypochondrialis species complex that represent the areas of relative climatic suitability over time. (k) The map of
the areas of stability represents the sum of the maps of all the periods analysed for the P. hypochondrialis species complex. The colour scales represent the suitability
(a)-(j) or stability (k) of the areas.

northwest. Some areas between these predicted refugia, as well as others concatenated phylogeny agree broadly with the previous hypotheses (e.
to the west and south, were predicted to be relatively stable, but not g., Faivovich et al., 2010; Bruschi et al., 2013; Duellman et al., 2016;
consistently by all the models (Fig. 5k). Haga et al., 2017; Andrade et al., 2020), that is, with the genus being
divided into two principal clades. The main differences are related to the
3.7. Spatiotemporal phylogeographic dynamics phylogenetic relationship of P. azureus and P. palliatus. The relationships
between P. palliatus and the other species of the genus have varied in
The most likely centre of diversification of the P. hypochondrialis previous studies, depending on the dataset or the methods used in the
species complex would have been located in present-day south-eastern phylogenetic reconstruction (Faivovich et al., 2010; Bruschi et al., 2013;
Amazonia (Fig. 6a - R), at the end of the Messinian age of the Miocene Duellman et al., 2016; Haga et al., 2017; Andrade et al., 2020). In the
(~5.87 Ma). During the Zanclean age of the Pliocene, the complex present phylogeny, P. palliatus was recovered as a sister of the campo
diversified and radiated northwards from this point through eastern rupestre and Atlantic Forest leaf frogs (i.e., P. ayeaye, P. centralis,
Amazonia and towards the southern limit between the Cerrado savanna P. megacephalus, P. oreades, P. rohdei, and P. rusticus). However, this
and the Amazon rainforest (Fig. 6b). During the Pleistocene, the lineages phylogenetic placement is only poorly supported (PP = 0.25), which
spread to the Llanos and the savannas of Guyana, the central Cerrado, highlights the need for more independent markers to better elucidate its
and the interface between eastern Amazonia and the Cerrado position in the arrangement.
(Fig. 6c–d). Overall, then, our results support an eastern Amazonian As in the present phylogeny, P. azureus has invariably been recovered
route during the Pliocene and rule out a possible Andean corridor during as the sister species of P. nordestinus in previous studies (Faivovich et al.,
the Pleistocene. The diffusion rate was rapid and constant, at a mean of 2010; Bruschi et al., 2013; Haga et al., 2017; Andrade et al., 2023),
~ 316.74 km/Ma (=0.316 m/year; 95 %-HPD = 0.189–0.486 m/year), except for Duellman et al.’s (2016) analysis, in which it was shown as a
with a subtle acceleration over the past two million years (Fig. 6e). sister of P. hypochondrialis. In this phylogeny, P. hypochondrialis was
represented by a chimera of six DNA fragments (mtDNA: 12S, 16S, and
4. Discussion cytb; nDNA: CXCR4, H3A, and NCX1) from distinct sources. Three of
these fragments were sequenced from the same individual, whose origin
4.1. Phylogenetic relationships was given only as “South America” and was purchased from an animal
dealer [GenBank accession numbers: AY948748 (16S), AY948826
The phylogenetic relationships of the Pithecopus species in the (NCX1), and FJ882741 (12S); Roelants et al., 2007; Van Bocxlaer et al.,

11
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Fig. 6. (a)-(d) Snapshots of the spatiotemporal dynamics of the Pithecopus hypochondrialis species complex in central-northern South America (the complete
reconstruction is available as a KML file in Supplementary File 1). R in circle (a) represents the estimated root location of the complex. The cyan polygons represent
the 80% highest posterior density (HPD) interval of uncertainty in the location of ancestral branches, and the lighter the tone, the older the ancestral branch
associated with it. The polygons with a black border (d) represent the current distribution of the lineages revealed by the first iteration of the general mixed Yule-
coalescent (GMYC) analysis (see Fig. 2, S3). (e) Diffusion rates over time, showing the mean values (solid line) and 95% HPD interval (dashed line). The green shape
shows the distribution of the Brazilian Cerrado savanna, as defined by the Brazilian Institute for Geography and Statistics Cerrado (IBGE, 2019), and WWF’s savanna
ecoregions (Olson et al., 2001), but excluding the Uruguayan savanna, and the dry and humid Chaco.

2009]. In a BLAST nucleotide search (Altschul et al., 1990), we found among these lineages, given that we were able to use only three inde­
that sequence AY948748 was 98.87–99.72 % similar to the available pendent loci for the phylogenetic analyses. Despite this, we were able to
sequences of P. azureus, while sequence FJ882741 was 99.42–99.56 % reassign specimens that had been mistakenly identified as
similar. In other words, the P. hypochondrialis sample analysed by P. hypochondrialis to P. araguaius and Pithecopus sp.
Duellman et al. (2016) is probably a chimera of P. azureus and
P. hypochondrialis, created by the misidentification of the purchased
frog, which would explain why the two species appear as sister taxa in 4.2. Species delimitation and taxonomic implications
this phylogeny. Finally, the phylogenetic relationship of Pithecopus sp.,
P. araguaius, and P. hypochondrialis were poorly supported in both the Our findings substantiate the conclusion that P. hypochondrialis
concatenated and the species trees, which emphasises the need for the constitutes a species complex which encompasses at least three
sequencing of additional independent loci to clarify the relationships morphologically cryptic species, including two that were previously
defined as P. araguaius and P. hypochondrialis. The third lineage may

12
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

represent a new taxon, which requires further evaluation and descrip­ 0.2 and 0.7, which is inconclusive for the confirmation of the evolu­
tion. Bruschi et al. (2013) obtained a tree with the same three-clade tionary independence of the lineages (Leaché et al., 2019). Overall, then,
layout. However, the samples in their subclade A and lineage 1 of the the evolutionary independence among the three validated species can
present study are not from the same localities, which prevents us from only be unequivocally assessed when we determine the best-fitting
determining whether they correspond to the same entity. Given this, we species tree and the θ and τ priors that best align with the observed
would recommend that any future studies assess whether the pop­ data. Given this, we would strongly recommend that any future phylo­
ulations from Belterra, in the Brazilian state of Pará, and Alta Floresta genetic investigations of this species complex incorporate more robust
(Mato Grosso) can belong to Pithecopus sp. If confirmed, the presence of data, such as genomic-scale markers or a more comprehensive matrix of
NORs in the short arm of the chromosome pair 8 (Bruschi et al., 2013) phenotypic traits.
emerges as a potential diagnosis for this candidate species. Our results confirm the taxonomic scenario outlined by Bruschi et al.
The results of the present study align with the more conservative (2013), which indicates that the P. hypochondrialis species complex is
hypothesis of the presence of three candidate species in the distributed throughout the Cerrado, Amazon, Llanos, and Guiana do­
P. hypochondrialis species complex. To reach this conclusion, we mains, whereas P. azureus is found in the Pantanal, Chaco, and Beni
implemented a two-iteration routine of the GMYC, given the evidence savannas, without no clear diagnosis based on colouration patterns.
that this increases accuracy when only a limited number of sequences is While this arrangement has been available for a decade, some studies
available for each species (Talavera et al., 2013). This approach resulted still adopt the classification of Caramaschi (2006), which considers the
in a considerable decrease in the number of candidate species for vali­ Cerrado populations to be P. azureus (e.g., Santoro & Brandão, 2014;
dation, which prevented overestimating the number of species delimi­ Thomassen & Ziade, 2020). This misconception may be especially
ted, as found in some previous studies that have used single-locus problematic in some contexts, such as morphological diagnoses and
delimitation methods to resolve under-sampled genetic variation population comparisons (Santos et al., 2018), and in particular in
(Lohse, 2009; Bergsten et al., 2012). It is important to note here that the ecological niche modelling (see Fig. 1 in Bandeira et al., 2021). Santos
sample analysed in the present study has a number of geographical gaps, et al. (2018), for example, reported variation in the formula of the tooth
in particular in the central portions of the Cerrado, Llanos and, Guiana row between individuals from the Argentinean Chaco [2(2)/2(1)] and
savannas, which may have led to the false identification of deep genetic the Brazilian Cerrado [2(2)/3(1)], and attributed this difference to
divergence between certain mitochondrial lineages. Given this, addi­ insufficient sampling in Argentina. However, the results of the present
tional geographic sampling will be essential to ascertain whether the study indicate that these populations are distinct species and that the
substantial genetic distances observed between some lineages, such as variation observed by Santos et al. (2018) would be diagnostic of these
2.1, 2.2, and 2.3 (Fig. 2; Appendix A), are the result of the actual species.
divergence of the lineages or rather, of sampling artefacts. Multi-species While colouration patterns have been used historically for the
coalescent delimitation methods also tend to overestimate the number diagnosis of phylomedusine taxa (e.g., Brandão, 2002; Caramaschi,
of candidate species when isolation-by-distance is strong, taxa are 2006; Giaretta et al., 2007), some authors have highlighted potential
widespread, or sampling is not representative of the actual species problems in the use of this type of trait, such as the pronounced intra­
ranges (Chambers and Hillis, 2020; Mason et al., 2020). Given that all specific variation and interspecific overlap (Baêta et al., 2009; Bruschi
three of these factors affect our dataset, the adoption of a three-species et al., 2013; Brunes et al., 2014; Magalhães et al., 2018). The results of
layout, rather than a 12-species scheme, would appear to be justifiably the present study are consistent with this, and reinforce the need for a
prudent, and avoid any gross overestimate of the number of valid taxonomic review of Pithecopus using other types of characters, such as
species. geometric morphometry, internal anatomy, larval morphology,
Conversely, we might ask – why were these lineages considered to be bioacoustics, cytogenetics, and behaviour.
a single species, rather than three? The null hypothesis of a single species
was not rejected in the second iteration of the ML-GMYC, although 4.3. Dynamics of range limits over time
several pieces of evidence contradict this conclusion, including the
result of the bGMYC, the ample phylogenetic distances between the The paleo-distribution and spatiotemporal reconstructions presented
clades (Fig. 2), and the clear gap between the intra- and inter-specific here reject the hypothesis of a Pleistocene corridor through the foothills
16S rDNA distances (Appendix A). The Pithecopus sp. specimens also of the Andes (Webb, 1991; Silva and Bates, 2002). On the other hand, an
diverged in their colouration pattern, given that this was the only line­ eastern Amazonian Pliocene route (Dowsett et al., 2016; Azevedo et al.,
age in which morphotype 3 predominated. In fact, while morphotype 3 2020) is supported by both approaches. The diversification of the
was present in 97 % of the Pithecopus sp. specimens, it was recorded in P. hypochondrialis species complex was estimated to have occurred be­
only 4.6 % of the P. hypochondrialis specimens, and was completely tween the Miocene and Pliocene, in south-eastern Amazonia. During the
absent in P. araguaius, further supporting our model-based hypothesis of Pliocene, this area was predicted to have had a highly suitable climate
the existence of three valid species. Haga et al. (2017) also distinguished for the distribution of the species complex. The dispersal of the common
P. araguaius from P. hypochondrialis based on their longer, more pulsated ancestor of the three lineages appears to have been restricted to this
calls, which are emitted at higher frequencies. While there was an region, which is consistent with the presence of savannas during this
overlap in the amplitude of 14 of the call traits analysed by these au­ period, as also predicted by Downset et al. (2016). To our knowledge,
thors, the differences could not explained by chance in four cases. only Leite et al. (2015) have provided any biological evidence for a
Ultimately, Solís-Lemus et al. (2015) concluded that the integrative Pliocene connection between the Cerrado and the Guiana Shield, spe­
iBPP approach increases the accuracy of the species delimitation by cifically for rodents, although they did not investigate or discuss the
limiting the parameter of space in which the models may under- or over- location or characteristics of this route.
estimate the number of candidate species, as well as reducing the po­ Our evidence of a central Amazonian or coastal Atlantic corridor for
tential impacts on the results of violations of the assumptions of the the P. hypochondrialis species complex is more uncertain, however. The
coalescent model. In the present study, most of the tested parameter paleo-distribution models projected highly suitable climates in eastern
combinations supported the existence of three species, even with the Amazonia and on the northern coast of South America throughout the
lack of clear morphometric differentiation. This further corroborates the Pleistocene, which may have favoured the establishment of a corridor in
three-species hypothesis. On the other hand, the results of the gdi were these regions. The continuous spatial reconstruction also indicates that
influenced considerably by the combination of priors used for the dis­ the Amazonian corridor was the most likely route connecting pop­
tribution of θ and τ, and the a priori species tree used in the iBPP ulations during this period. Despite proposals of a central Amazonian
delimitations. In most input combinations, the gdi values ranged from corridor (Silva and Bates, 2002), the sum of the ecological modelling

13
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

evidence – including that from the present study – indicates that the 4.4. Conclusions and perspectives
most likely location of this route would have been in eastern Amazonia
(e.g., Bueno et al., 2017; Buzatti et al., 2017; Buzatti et al., 2018). Ox­ We have presented an extensive and integrated analysis of the mo­
ygen isotope data from speleothems also demonstrate that eastern lecular, morphological, and ecological data, which offers valuable in­
Amazonia has been drier since at least the last two glacial-interglacial sights into the diversification of the P. hypochondrialis species complex.
cycles (Cheng et al., 2013), although pollen fossils indicate that the It is nevertheless important to acknowledge the limitations of the sam­
tropical evergreen forests were replaced by seasonal forests rather than ple, in particular in the northern extreme of the clade’s geographic
being replaced by savanna or grassland (Kern et al., 2023). However, the distribution and narrow genomic sampling. The prevalence of cryptic
extent to which these changes affected the connectivity of savanna- diversity in the genus Pithecopus has been revealed in several previous
associated species in South America remains unclear, and the general studies of molecular markers (Faivovich et al., 2010; Ramos et al., 2019;
paucity of fossils continues to be the principal challenge for the defini­ Andrade et al., 2020). In line with these studies, our findings also
tive determination of if, when, and how this corridor may have been indicate the existence of at least one undescribed species in the
formed (Azevedo et al., 2020). P. hypochondrialis species complex. It is interesting to note that only
In the present study, sampling is especially deficient across the Haga et al. (2017) and Andrade et al. (2020) described formally the
northern portion of the species’ distribution, with very poor represen­ cryptic species detected in their molecular studies. Haga et al. (2017)
tation from Colombia, French Guiana, Guyana, and Suriname, and no described P. araguaius, which is indistinguishable from P. hypochon
sequences from Venezuelan specimens. As continuous phylogeographic drialis, based on its morphology and bioacoustics, while Andrade et al.
methods are known to be affected strongly by sampling gaps (Kalkaus­ (2020) described P. gonzagai, which was similarly indistinguishable
kas et al., 2021), we suspect that the apparent long-distance jump in from P. nordestinus. Even so, we do not rule out altogether the possibility
dispersal from the Cerrado to the Llanos through central Amazonia is that these supposedly cryptic species and Pithecopus sp. may become
almost certainly a sampling artefact. Given this, a more comprehensive distinguishable when analysed through other character systems. Given
sample of the lineages delimited in the present study, in particular from the ongoing global biodiversity crisis and widespread Linnaean short­
the areas with greatest data deficiencies, should provide a more realistic falls (Hortal et al., 2015), it will be essential to review and describe the
overview of the history of the geographical distribution of the species cryptic species identified within the genus Pithecopus. This will permit
complex. At this point, the only Pleistocene route that can be discarded their inclusion in future assessments of threat risks and the development
with confidence is the Andean corridor. A more comprehensive sample, of conservation policies. In particular, future research efforts on Pith­
especially from the northern extreme of its range, will be crucial to ecopus should focus on the existing geographic gaps, the expansion of
determine whether the coastal, eastern Amazonian, or a combination of genomic sampling, and the integration of phenotypic data from multiple
both routes may account for the connectivity of the P. hypochondrialis sources. These additional perspectives will contribute to a more
species complex in the Pleistocene savannas of South America. comprehensive understanding of the characteristics and geographic
The mean diffusion rate of the P. hypochondrialis species complex was distribution of these species.
considerably higher than those estimated previously for other Although our results offer mixed support for the eastern Amazonian
Neotropical anurans. Bokermannohyla saxicola has a mean diffusion rate and coastal Atlantic Pleistocene corridors, our evidence clearly supports
approximately 4.5 times lower than the rate we estimated here (Oswald the occurrence of pulses of connectivity between the northern and
et al., 2022), for example, while the mean diffusion rates of P. ayeaye southern Neotropical savannas throughout both the Pliocene and the
and P. megacephalus were around 2.3 times lower, with a comparable Pleistocene. Previous studies have already presented biological evidence
rate only being observed during a period of expansion followed by an of the connectivity of organisms that inhabited the open areas of the
introgression event that occurred between the two species at approxi­ Brazilian and Guyana shields during the Pliocene (Leite et al., 2015).
mately 300 ka (Magalhães et al., 2021). However, B. saxicola, P. ayeaye, However, our findings represent the first phylogeographic evidence of
and P. megacephalus all have considerably more restricted geographic the existence of a corridor in eastern Amazonia during the late Pliocene.
ranges than the P. hypochondrialis species complex, given that they are Although we have assessed the spatiotemporal dynamics of the
endemic to highland campo rupestre and are specialised in the use of P. hypochondrialis species complex, a number of other phylogeographic
rocky environments for reproduction (Magalhães et al., 2021; Oswald questions remain unanswered, such as how Plio-Pleistocene climate
et al., 2022). changes affected the demographic history and dynamics of the gene flow
Even so, the mean diffusion rate predicted here for P. hypochondrialis within the group. Given its ample distribution across multiple
is only approximately a third of that estimated for Pseudopaludicola Neotropical domains and its occurrence within the interface of open and
falcipes by Langone et al. (2016). This leptodactylid frog has an extensive forested habitats, the P. hypochondrialis species complex appears to be a
geographical distribution and thrives in the Pampa grasslands of promising model for the investigation of the environmental dynamics of
southern South America, where it is relatively abundant. A high diffu­ the Neotropics.
sion rate was projected for this species because of its presumed mech­
anisms of passive dispersal, such as basin captures and vegetation rafts CRediT authorship contribution statement
(Langone et al., 2016). These assumptions are derived from the obser­
vation that these frogs vocalise in waterlogged areas adjacent to streams, Rafael F. Magalhães: Conceptualization, Investigation, Formal
environments that are conductive to dispersal events of this type. The analysis, Resources, Writing – original draft. Elisa K. S. Ramos: Inves­
passive dispersal hypothesis is further supported by the diminutive size tigation, Formal analysis, Writing – review & editing. Lucas N. Ban­
of Ps. falcipes, which would limit considerably its potential for active deira: Investigation, Formal analysis, Writing – review & editing.
long-distance dispersal. By contrast, the members of the Johnny S. Ferreira: Data curation, Writing – review & editing. Fer­
P. hypochondrialis species complex are neither miniaturised nor do they nanda P. Werneck: Conceptualization, Supervision, Writing – review &
breed in riparian environments, which contradicts any hypothesis of editing. Marina Anciães: Conceptualization, Supervision, Writing –
passive dispersal. In this case, the high diffusion rate observed in this review & editing. Daniel P. Bruschi: Conceptualization, Data curation,
group is likely explained by its generalist habits (Barrio-Amorós, 2009) Resources, Supervision, Funding acquisition, Writing – review & editing.
and its environmental tolerance, which would underpin a much greater
potential for active dispersal and the efficient colonisation of new Declaration of Competing Interest
regions.
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence

14
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

the work reported in this paper. Funding

Acknowledgements This study was supported by Minas Gerais State Research Founda­
tion, FAPEMIG (grant APQ00325-21), the São Paulo State Research
We thank all the museum and collection curators, in particular, Célio Foundation, FAPESP (grant 2012/12169-7), the Amazonas State
F. B. Haddad, Luís F. Toledo and Natan M. Maciel, for making the ma­ Research Foundation, FAPEAM (grant 01.02.016301.03263/2021-47),
terial under their care available to us for analysis. We also thank Fausto the Brazilian National Council for Scientific and Technological Devel­
Nomura for helping with the verification of the material. We would opment, CNPq (grants 425571/2018-1, 303646/2021-7), the Rufford
especially like to thank Shirlei Maria Recco-Pimentel for her guidance Foundation (grants 23763-1, 27629-2); DPB was funded by a FAPESP
during the initial study, and her help with its design and the technical scholarship (2010/17464-1) and CNPq (grant 303646/2021-7). LNB
and financial infrastructure, and Luciana Bolsoni Lourenço for her was supported by a CAPES fellowship (grant 001), EKSR was funded by a
valuable comments during the research and on a previous version FAPESP doctoral scholarship (grant 2018/01236-1), and; FPW was
of the manuscript. We are also very grateful to Melisa Olave and three supported by a CNPq productivity fellowship (grant 311504/2020-5).
anonymous reviewers for their valuable contributions that helped to
improve this paper greatly.

Appendix A

Mean uncorrected p-distances recorded between the 16S sequences. Intra-lineage distances are shown in the diagonal (in bold type) and the inter-
lineage distances are given in the lower half of the table. The individuals of the lineages resulting from the first and second iterations of the GMYC are
shown in Fig. 2. NA = not applicable.

Lineage 1st iteration of GMYC

1.1 1.2 1.3 2.1 2.2 2.3 3.1 3.2 3.3 3.4 3.5 3.6

1.1 0.001
1.2 0.025 NA
1.3 0.023 0.012 0.003
2.1 0.038 0.024 0.030 0.002
2.2 0.051 0.044 0.045 0.031 NA
2.3 0.052 0.044 0.046 0.025 0.037 0.009
3.1 0.049 0.042 0.040 0.024 0.049 0.044 0.004
3.2 0.049 0.046 0.046 0.035 0.056 0.052 0.023 0.006
3.3 0.048 0.043 0.042 0.031 0.051 0.047 0.017 0.026 0.009
3.4 0.046 0.043 0.041 0.031 0.052 0.047 0.019 0.024 0.009 0.002
3.5 0.051 0.049 0.048 0.036 0.055 0.051 0.022 0.029 0.015 0.014 0.007
3.6 0.044 0.041 0.040 0.027 0.051 0.047 0.020 0.023 0.013 0.011 0.016 0.006

Lineage 2nd iteration of GMYC


1 2 3
1 0.015
2 0.043 0.023
3 0.045 0.042 0.017

Appendix B. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ympev.2023.107959.

References Avila-Pires, T.C.S., 1995. Lizards of Brazilian Amazonia (Reptilia: Squamata). Zool. Verh.
Leiden 299, 1–706. https://repositorio.museu-goeldi.br/handle/mgoeldi/1222.
Azevedo, J.A.R., Collevatti, R.G., Jaramillo, C.A., Strömberg, C.A.E., Guedes, T.B., Matos-
Abreu-Jardim, T.P.F., Lima, N.E., Jardim, L., Maciel, N.M., Magalhães, R.F., Colli, G.R.,
Marví, P., Bacon, C.D., Carillo, J.D., Faurby, S., Antonelli, A., 2020. On the young
Haddad, C.F.B., Collevatti, R.G., 2023. Neogene-Quaternary tectonic, eustatic and
savannas in the land of ancient forest. In: Rull, V., Carnaval, A.C. (Eds.), Neotropical
climatic events shaped the evolution of a South American treefrog. J. Biogeogr. 50,
Diversification: Patterns and Processes. Springer, Cham, Switzerland, pp. 271–298.
987–999. https://doi.org/10.1111/jbi.14578.
https://doi.org/10.1007/978-3-030-31167-4_12.
Akinwande, M., Dikko, H., Samson, A., 2015. Variance Inflation Factor: As a Condition
Baêta, D., Caramaschi, U., Cruz, C.A.G., Pombal Jr., J.P., 2009. Phyllomedusa itacolomi
for the Inclusion of Suppressor Variable(s) in Regression Analysis. Open J. Stat. 5,
Caramaschi, Cruz & Feio, 2006, as junior synonym of Phyllomedusa ayeaye (B. Lutz,
754–767. https://doi.org/10.4236/ojs.2015.57075.
1966) (Hylidae, Phyllomedusinae). Zootaxa 2226, 58–65. 10.11646/zootaxa.2226
Aktas, C., 2020. haplotypes: manipulating DNA sequences and estimating unambiguous
.1.5.
haplotype network with statistical parsimony. Available at R Package Version 1 (1),
Bandeira, L.N., Villalobos, F., Werneck, F.P., Townsend Peterson, A., Anciães, M., 2021.
2. https://CRAN.R-project.org/package=haplotypes.
Different elevational environments dictate contrasting patterns of niche evolution in
Altschul, S.F., Gish, W., Miller, W., Myers, E.W., Lipman, D.J., 1990. Basic local
Neotropical Pithecopus treefrog species. Biotropica 53, 1042–1051. https://doi.org/
alignment search tool. J. Mol. Biol. 215, 403–410. https://doi.org/10.1016/S0022-
10.1111/btp.12929.
2836(05)80360-2.
Barrio-Amorós, C.L., 2009. Distribución y aspectos de la historia natural de las ranas
Andrade, F.S., Haga, I.A., Ferreira, J.S., Recco-Pimentel, S.M., Toledo, L.F., Bruschi, D.P.,
lémur (Hylidae: Phyllomedusinae) en Venezuela. Mem. Fund. La Salle Cienc. Nat.
2020. A new cryptic species of Pithecopus (Anura, Phyllomedusidae) in north-
171, 19–46.
eastern Brazil. Eur. J. Taxon. 723, 108–134. https://doi.org/10.5852/
Bergsten, J., Bilton, D.T., Fujisawa, T., Elliott, M., Monaghan, M.T., Balke, M.,
ejt.2020.723.1147.
Hendrich, L., Geijer, J., Herrmann, J., Foster, G.N., Ribera, I., Nilsson, A.N.,

15
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Barraclough, T.G., Vogler, A.P., 2012. The effect of geographical scale of sampling Duellman, W.E., 1970. The Hylid Frogs of Middle America, first ed. University of Kansas,
on DNA barcoding. Syst. Biol. 61, 851–869. https://doi.org/10.1093/sysbio/sys037. Lawrence, Kansas. 10.5962/bhl.title.2835.
Bernarde, P.S., 2007. Calling period and environments of anuran fauna in municipality of Elith, J., Phillips, S.J., Hastie, T., Dudík, M., Chee, Y.E., Yates, C.J., 2011. A statistical
Espigão do Oeste, Rondônia, Southeastern of Amazonian – Brazil (Amphibia: Anura). explanation of MaxEnt for ecologists. Divers. Distrib. 17, 43–57. https://doi.org/
Biota Neotrop. 7, 88–92. https://doi.org/10.1590/S1676-06032007000200010. 10.1111/j.1472-4642.2010.00725.x.
Bielejec, R., Rambaut, A., Suchard, M.A., Lemey, P., 2011. SPREAD: spatial phylogenetic Excoffier, L., Lischer, H.E.L., 2010. Arlequin suite ver 3.5: A new series of programs to
reconstruction of evolutionary dynamics. Bioinformatics 27, 2910–2912. https:// perform population genetics analyses under Linux and Windows. Mol. Ecol. Res. 10,
doi.org/10.1093/bioinformatics/btr481. 564–567. https://doi.org/10.1111/j.1755-0998.2010.02847.x.
Bouckaert, R.R., Drummond, A.J., 2017. bModelTest: Bayesian phylogenetic site model Faivovich, J., Haddad, C.F.B., Baêta, D., Jungfer., Álvares, G.F.R, Brandão, R.A., Sheil, C.,
averaging and model comparison. BMC Evol. Biol. 17, 42. https://doi.org/10.1186/ Barrientos, L.S., Barrio-Amorós, C.L., Cruz, C.A.G., Wheeler, W.C., 2010. The
s12862-017-0890-6. phylogenetic relationships of the charismatic poster frogs, Phyllomedusinae (Anura,
Bouckaert, R.R., Vaughan, T.G., Barido-Sottani, J., Duchêne, S., Fourment, M., Hylidae). Cladistics 26, 227–261. 10.1111/j.1096-0031.2009.00287.x.
Gavryushkina, A., Heled, J., Jones, G., Kühnert, D., de Maio, N., Matschiner, M., Flot, J.F., 2010. SeqPHASE: a web tool for interconverting phase input/output files and
Mendes, F.K., Müller, N.F., Ogilvie, H.A., du Plessis, L., Popinga, A., Rambaut, A., fsata sequence alignments. Mol. Ecol. Resour. 10, 162–166. https://doi.org/
Rasmussen, D., Siveroni, I., Suchard, M.A., Wu, C., Xie, D., Zhang, C., Stadler, T., 10.1111/j.1755-0998.2009.02732.x.
Drummond, A.J., 2019. BEAST 2.5: an advanced software platform for Bayesian Fonte, L.F.M., Mayer, M., Lötters, S., 2019. Long-Distance Dispersal in Amphibians.
evolutionary analysis. PLoS Comput. Biol. 15, e1006650. Front. Biogeogr. 11, e44577. 10.21425/F5FBG44577.
Brandão, R.A., 2002. A new species of Phyllomedusa Wagler, 1830 (Anura: Hylidae) Fordham, D.A., Saltré, F., Haythorne, S., Wigley, T.M.L., Otto-Bliesner, B.L., Chan, K.C.,
from Central Brazil. J. Herp. 36, 571–578. https://doi.org/10.1670/0022-1511 Brook, B.W., 2017. PaleoView: a tool for generating continuous climatic projections
(2002)036[0571:ANSOPW]2.0.CO;2. spanning the last 21 000 years at regional and global scales. Ecography 40,
Brown, J.L., Hill, D.J., Dolan, A.M., Carnaval, A.C., Haywood, A.M., 2018. Data 1348–1358. https://doi.org/10.1111/ecog.03031.
Descriptor: PaleoClim, high spatial resolution paleoclimate surfaces for global land Fox, J., Weisberg, S., 2018. An R companion to applied regression, third ed. Sage
areas. Sci. Data 5, 180254 (2018). 10.1038/sdata.2018.254. Publications, Thousand Oaks, California.
Brunes, T.O., Alexandrino, J., Baêta, D., Zina, J., Haddad, C.F.B., Sequeira, F., 2014. Frost, D.R., 2023. Amphibian Species of the World: an online reference. Version 6.1. Am.
Species limits, phylogeographic and hybridization patterns in Neotropical leaf frogs Mus. Nat. Hist. https://amphibiansoftheworld.amnh.org/index.php (accessed 9
(Phyllomedusinae). Zool. Scr. 43, 586–604. https://doi.org/10.1111/zsc.12079. March 2023). 10.5531/db.vz.0001.
Bruschi, D.P., Busin, C.S., Siqueira, S., Recco-Pimentel, S.M., 2012. Cytogenetic analysis Gehara, M., Crawford, A.J., Orrico, V.G.D., Rodríguez, A., Lötters, S., Fouquet, A.,
of two species in the Phyllomedusa hypochondrialis group. Hereditas 149, 34–40. Barrientos, L.S., Brusquetti, F., de la Riva, I., Ernst, R., Urrutia, G.G., Glaw, F.,
https://doi.org/10.1111/j.1601-5223.2010.02236.x. Guayasamin, J.M., Hölting, M., Jansen, M., Kok, P.J.R., Kwet, A., Lingnau, R.,
Bruschi, D.P., Busin, C.S., Toledo, L.F., Vasconcellos, G.A., Strüssmann, C., Weber, L.N., Lyra, M., Moravec, J., Pombal Jr., J.P., Rojas-Runjaic, F.J.M., Schulze, A., Señaris, J.
Lima, A.P., Lima, J.D., Recco-Pimentel, S.M., 2013. Evaluation of the taxonomic C., Solé, M., Rodrigues, M.T., Twomey, E., Haddad, C.F.B., Vences, M., Köhler, J.,
status of populations assigned to Phyllomedusa hypochondrialis (Anura, Hylidae, 2014. High levels of diversity uncovered in a widespread nominal taxon: continental
Phyllomedusinae) based on molecular, chromosomal, and morphological approach. phylogeography of the Neotropical tree frog Dendropsophus minutus. PLoS ONE 9,
BMC Genet. 14, 70. https://doi.org/10.1186/1471-2156-14-70. e103958.
Bueno, M.L., Pennington, R.T., Dexter, K.G., Kamino, L.H.Y., Pontara, V., Neves, D.M., Gernhard, T., 2008. The conditioned reconstructed process. J. Theor. Biol. 253, 769–778.
Ratter, J.A., Oliveira-Filho, A.T., 2017. Effects of Quanternary climatic fluctuations https://doi.org/10.1016/j.jtbi.2008.04.005.
on the distribution of Neotropical savanna tree species. Ecography 39, 1–12. https:// Giaretta, A.A., Oliveira Filho, J.C., Kokubum, M.N.C., 2007. A new Phyllomedusa Wagler
doi.org/10.1111/ecog.01860. (Anura, Hylidae) with reticulated pattern on flanks from Southeastern Brazil.
Buzatti, R.S.O., Lemos-Filho, J.P., Bueno, M.L., Lovato, M.B., 2017. Multiple Pleistocene Zootaxa 1614, 31–41. 10.11646/zootaxa.1614.1.3.
refugia in the Brazilian cerrado: evidence from phylogeography and climatic niche Guarnizo, C.E., Cannatella, D.C., 2013. Genetic divergence within frog species is greater
modelling of two Qualea species (Vochysiaceae). Bot. J. Linn. Soc. 185, 307–320. in topographically more complex regions. J. Zoolog. Syst. Evol. Res. 51, 333–340.
https://doi.org/10.1093/botlinnean/box062. https://doi.org/10.1111/jzs.12027.
Buzatti, R.S.O., Pfeilsticker, T.R., Magalhães, R.F., Bueno, M.L., Lemos-Filho, J.P., Guarnizo, C.E., Paz, A., Muñoz-Ortiz, A., Flechas, S.V., Méndez-Narváez, J., Crawford, A.
Lovato, M.B., 2018. Genetic and historical colonization analyses of an endemic J., 2015. DNA barcoding survey of anurans across the eastern cordillera of the Andes
savanna tree, Qualea grandiflora, reveal ancient connections between amazonian on cryptic diversity. PLoS ONE 10, e0127312.
savannas and Cerrado core. Front. Plant Sci. 9, 981. https://doi.org/10.3389/ Guillot, G., Mortier, F., Estoup, A., 2005. GENELAND: a computer package for landscape
fpls.2018.00981. genetics. Mol. Ecol. Notes 5, 712–715. https://doi.org/10.1111/j.1471-
Caramaschi, U., 2006. Redefinição do grupo de Phyllomedusa hypochondrialis, com 8286.2005.01031.x.
redescrição de P. megacephala (Miranda-Ribeiro, 1926), revalidação de P. azurea Guillot, G., Renaud, S., Ledevin, R., Michaux, J., Claude, J., 2012. A unifying model for
Cope, 1862 e descrição de uma nova espécie (Amphibia, Anura, Hylidae). Arq. Mus. the analysis of phenotypic, genetic, and geographic data. Syst. Biol. 61, 897–911.
Nac. (rio De J.) 64, 159–179. https://biostor.org/reference/249183. https://doi.org/10.1093/sysbio/sys038/.
Carnaval, A.C., Moritz, C., 2008. Historical climate modelling predicts patterns of current Haga, I.A., Andrade, F.S., Bruschi, D.P., Recco-Pimentel, S.M., Giaretta, A.A., 2017.
biodiversity in the Brazilian Atlantic forest. J. Biogeogr. 35, 1187–1201. https://doi. Unrevealing the leaf frogs Cerrado diversity: a new species of Pithecopus (Anura,
org/10.1111/j.1365-2699.2007.01870.x. Arboranae, Phyllomedusidae) from Mato Grosso state. Brazil. Plos ONE 12,
Carstens, B.C., Pelletier, T.A., Reid, N.M., Satler, J.D., 2013. How to fail at species e0184631.
delimitation. Mol. Ecol. 22, 4369–4383. https://doi.org/10.1111/mec.12413. Heyer, W.R., Rand, A.S., da Cruz, C.A.G., Peixoto, O.L., Nelson, C.E., 1990. Frogs of
Chambers, E.A., Hillis, D.M., 2020. The multispecies coalescent over-splits species in the Boraceia. Arq. Zool. (são Paulo) 31, 231–410. 10.11606/issn.2176-7793.v31i4p231
case of geographically widespread taxa. Syst. Biol. 69, 184–193. https://doi.org/ -410.
10.1093/sysbio/syz042. Hill, D.J., 2015. The non-analogue nature of Pliocene temperature gradients. Earth
Cheng, H., Sinha, A., Cruz, F.W., Wang, X., Edwards, R.L., d’Horta, F.M., Ribas, C.C., Planet. Sci. Lett. 425, 232–241. https://doi.org/10.1016/j.epsl.2015.05.044.
Vuille, M., Stott, L.D., Auler, A.S., 2013. Climate change patterns in Amazonia and IBGE (Instituto Brasileiro de Geografia e Estatística), 2019. Biomas e sistema costeiro-
biodiversity. Nat. Commun. 4, 1411. https://doi.org/10.1038/ncomms2415. marinho do Brasil: compatível com a escala 1:250000. IBGE, Coordenação de
Cobos, M.E., Peterson, A.T., Barve, N., Osorio-Olvera, L., 2019. kuenm: an R package for Recursos Naturais e Estudos Ambientais, Rio de Janeiro, Brazil.
detailed development of ecological niche models using Maxent. PeerJ 7, e6281. Jackson, N.D., Carstens, B.C., Morales, A.E., O’Meara, B.C., 2017. Species delimitation
Crawford, A.J., 2003. Huge populations and old species of Costa Rican and Panamanian with gene flow. Syst. Biol. 66, 799–812. https://doi.org/10.1093/sysbio/syw117.
dirt frogs inferred from mitochondrial and nuclear gene sequencies. Mol. Ecol. 12, Jones, G., 2017. Algorithmic improvements to species delimitation and phylogeny
2525–2540. https://doi.org/10.1046/j.1365-294X.2003.01910.x. estimation under the multispecies coalescent. J. Math. Biol. 74, 447–467. https://
Dolan, A.M., Haywood, A.M., Hunter, S.J., Tindall, J.C., Dowsett, H.J., Hill, D.J., doi.org/10.1007/s00285-016-1034-0.
Pickering, S.J., 2015. Modelling the enigmatic Late Pliocene Glacial Event – Marine Kalkauskas, A., Perron, U., Sun, Y., Goldman, N., Baele, G., Guindon, S., De Maio, N.,
Isotope Stage M2. Glob. Planet. Change 128, 47–60. https://doi.org/10.1016/j. 2021. Sampling bias and model choice in continuous phylogeography: getting lost on
gloplacha.2015.02.001. random walk. PLoS Comput. Biol. 17, e1008561.
Dowsett, H., Dolan, A., Rowley, D., Moucha, R., Forte, A.M., Mitrovica, J.X., Pound, M., Katoh, K., Rozewicki, J., Yamada, K.D., 2019. MAFFT online service: multiple sequence
Salzmann, U., Robinson, M., Chandler, M., Foley, K., Haywood, A., 2016. The alignment, interactive sequence choice and visualization. Brief. Bioinform. 20,
PRISM4 (mid-Piacenzian) paleoenvironmental reconstruction. Clim. past. 12, 1160–1166. https://doi.org/10.1093/bib/bbx108.
1519–1538. https://doi.org/10.5194/cp-12-1519-2016. Katoh, K., Standley, D.M., 2013. MAFFT multiple sequence alignment software version 7:
Drummond, A.J., Ho, S.Y.W., Phillips, M.J., Rambaut, A., 2006. Relaxed phylogenetics improvements in performance and usability. Mol. Biol. Evol. 30, 772–780. https://
and dating with confidence. PLoS Biol. 4, e88. doi.org/10.1093/molbev/mst010.
Drummond, A.J., Suchard, M.A., Xie, D., Rambaut, A., 2012. Bayesian phylogenetics Katoh, K., Toh, H., 2008. Recent developments in the MAFFT multiple sequence
with BEAUti and the BEAST 1.7. Mol. Biol. Evol. 29, 1969–1973. https://doi.org/ alignment program. Brief. Bioinform. 9, 286–298. https://doi.org/10.1093/bib/
10.1093/molbev/mss075. bbn013.
Duellman, W.E., Trueb, L., 1994. Biology of Amphibians, first ed. The John Hopkins Kern, A.K., Akabane, T.K., Ferreira, J.Q., Chiessi, C.M., Willard, D.A., Ferreira, F.,
University Press, Baltimore, Maryland. Sanders, A.O., Silva, C.G., Rigsby, C., Cruz, F.W., Dwyer, G.S., Fritz, S.C., Baker, P.A.,
Duellman, W.E., Marion, A.B., Hedges, S.B., 2016. Phylogenetics, classification, and 2023. A 1.8 million year history of Amazon vegetation. Quat. Sci. Rev. 299, 107867
biogeography of the treefrogs (Amphibia: Anura: Arboranae). Zootaxa 4104, 1–109. https://doi.org/10.1016/j.quascirev.2022.107867.
10.11646/zootaxa.4104.1.1.

16
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Kumar, S., Stecher, G., Li, M., Knyaz, C., Tamura, K., 2018. MEGA X: molecular Posada, D., Crandall, K.A., Templeton, A.R., 2000. GeoDis: a program for the cladistic
evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 35, nested analysis of the geographical distribution of genetic haplotypes. Mol. Ecol. 9,
1547–1549. https://doi.org/10.1093/molbev/msy096. 487–488. https://doi.org/10.1046/j.1365-294x.2000.00887.x.
Langone, J.A., Camargo, A., de Sá, R.O., 2016. High genetic diversity but low population Prado, C.P.A., Haddad, C.F.B., Zamudio, K.R., 2012. Cryptic lineages and Pleistocene
structure in the frog Pseudopaludicola falcipes (Hensel, 1867) (Amphibia, Anura) population expansion in a Brazilian Cerrado frog. Mol. Ecol. 21, 921–941. https://
from the Pampas of South America. Mol. Phylogenet. Evol. 95, 137–151. doi.org/10.1111/j.1365-294X.2011.05409.x.
Leaché, A.D., Zhu, T., Rannala, B., Yang, Z., 2019. The spectre of too many species. Syst. Rambaut, A., Drummond, A.J., Xie, D., Baele, G., Suchard, M.A., 2018. Posterior
Biol. 68, 168–181. https://doi.org/10.1093/sysbio/syy051. summarization in Bayesian phylogenetics using Tracer 1.7. Syst. Biol. 67, 901–904.
Leite, Y.L.R., Kok, P.J.R., Weksler, M., 2015. Evolutionary affinities of the ‘Lost World’ https://doi.org/10.1093/sysbio/syy032.
mouse suggest a late Pliocene connection between the Guiana and Brazilian shields. Ramos, E.K.S., Magalhães, R.F., Marques, N.C.S., Baêta, D., Garcia, P.C.A., Santos, F.R.,
J. Biogeogr. 42, 706–715. https://doi.org/10.1111/jbi.12461. 2019. Cryptic diversity in Brazilian endemic monkey frogs (Hylidae,
Lemey, P., Rambaut, A., Welch, J.J., Suchard, M.A., 2010. Phylogeography takes a Phyllomedusinae, Pithecopus) revealed by multispecies coalescent and integrative
relaxed random walk in continuous space and time. Mol. Biol. Evol. 27, 1877–1885. approaches. Mol. Phylogenet. Evol. 132, 105–116. https://doi.org/10.1016/j.
https://doi.org/10.1093/molbev/msq067. ympev.2018.11.022.
Lescure, J., Marty, V.C., Starace, F., Auber-Thomay, M., Letellier, F., 1995. Contribution Reid, N.M., Carstens, B.C., 2012. Phylogenetic estimation error can decrease the
à l’étude des amphibiens de Guyane française. X. Les Phyllomedusa (Anura, accuracy of species delimitation: a Bayesian implementation of the general mixed
Hylidae). Revue Fr. Aquariol. 22, 35–50. Yule-coalescent model. BMC Evol. Biol. 12, 196. https://doi.org/10.1186/1471-
Lima, J.R.F., Lima, J.D., Lima, S.D., Silva, R.B.L., Andrade, G.V., 2017. Amphibians found 2148-12-196.
in the Amazonian Savanna of the Rio Curiaú environmental protection area in Rodríguez, A., Börner, M., Pabijan, M., Gehara, M., Haddad, C.F.B., Vences, M., 2015.
Amapá. Brazil. Biota Neotrop. 17, e20160252. Genetic divergence in tropical anurans: deeper phylogeographic structure in forest
Lima-Rezende, C.A., Souza, R.O., Caparroz, R., 2019. The spatial genetic structure of the specialists and in topographically complex regions. Evol. Ecol. 29, 765–785. https://
white-banded tanager (Aves, Passeriformes) in fragmented Neotropical savannas doi.org/10.1007/s10682-015-9774-7.
suggests two evolutionarily significant units. Biotropica 51, 234–244. https://doi. Roelants, K., Gower, D.J., Wilkinson, M., Loader, S.P., Biju, S.D., Guillaume, K.,
org/10.1111/btp.12623. Moriau, L., Bossuyt, F., 2007. Global patterns of diversification in the history of
Liu, C., White, M., Newell, G., 2013. Selecting thresholds for the prediction of species modern amphibians. PNAS 104, 887–892. https://doi.org/10.1073/
occurrence with presence-only data. J. Biogeogr. 40, 778–789. https://doi.org/ pnas.0608378104.
10.1111/jbi.12058. Rosenberg, N.A., 2004. DISTRUCT: a program for the graphical display of population
Lleonart, J., Salat, J., Torres, G.J., 2000. Removing allometric effects of body size in structure. Mol. Ecol. Notes 4, 137–138. https://doi.org/10.1046/j.1471-
morphological analysis. J. Theor. Biol. 205, 85–93. https://doi.org/10.1006/ 8286.2003.00566.x.
jtbi.2000.2043. Salzburger, W., Ewing, G.B., Von Haeseler, A., 2011. The performance of phylogenetic
Lohse, K., 2009. Can mtDNA barcodes be used to delimit species? A response to Pons algorithms in estimating haplotype genealogies with migration. Mol. Ecol. 20,
et al. (2006). Syst. Biol. 58, 439–442. 10.1093/sysbio/syp039. 1952–1963. https://doi.org/10.1111/j.1365-294X.2011.05066.x.
Magalhães, R.F., Rocha, P.C., Santos, F.R., Strüssmann, C., Giaretta, A.A., 2018. Santoro, G.R.C.C., Brandão, R.A., 2014. Reproductive modes, habitat use, and richness of
Integrative taxonomy helps to assess the extinction risk of anuran species. J. Nat. anurans from Chapada dos Veadeiros, central Brazil. North-West. J. Zool. 10,
Conserv. 45, 1–10. https://doi.org/10.1016/j.jnc.2018.07.001. 365–373.
Magalhães, R.F., Lemes, P., Santos, M.T.T., Mol, R.M., Ramos, E.K.S., Oswald, C.B., Santos, D.L., Morais, A.R., Signorelli, L., Bastos, R.P., Feio, R., Nomura, F., 2018.
Pezzuti, T.L., Santos, F.R., Brandão, R.A., Garcia, P.C.A., 2021. Evidence of Description of the tadpole of Phyllomedusa azurea from the Brazilian Cerrado, with a
introgression in endemic frogs from the campo rupestre contradicts the reduced description of the internal oral morphology of Phyllomedusa oreades. Herpetologica
hybridization hypothesis. Biol. J. Linn. Soc. 133, 561–576. https://doi.org/10.1093/ 74, 50–57. https://doi.org/10.1655/HERPETOLOGICA-D-16-00050.1.
biolinnean/blaa142. Schmid, M., Steinlein, C., Feichtinger, W., Bogart, J.P., 2014. Chromosome banding in
Mantel, N., 1967. The detection of disease clustering and a generalized regression Amphibia. XXXI. The Neotropical anuran families Centrolenidae and Allophrynidae.
approach. Cancer Res. 27, 209–220. Cytogenet. Genome Res. 142, 268–285. https://doi.org/10.1159/000362216.
Martins, L.F., Choueri, E.L., Oliveira, A.F.S., Domingos, F.M.C.B., Caetano, G.H.O., Silva, J.M.C., Bates, J.M., 2002. Biogeographic patterns and conservation in the South
Cavalcante, V.H.G.L., Leite, R.N., Fouquet, A., Rodrigues, M.T., Carnaval, A.C., American Cerrado: a tropical savanna hotspot. BioScience 52, 225–233. https://doi.
Colli, G.R., Werneck, F.P., 2021. Whiptail lizard lineage delimitation and population org/10.1641/0006-3568(2002)052[0225:BPACIT]2.0.CO;2.
expansion as windows into the history of Amazonian open ecosystems. Syst. Biod. Simpson, G.G., 1951. The species concept. Evolution 5, 285–298. https://doi.org/
19, 957–975. https://doi.org/10.1080/14772000.2021.1953185. 10.1111/j.1558-5646.1951.tb02788.x.
Mason, N.A., Fletcher, N.K., Gill, B.A., Funk, W.C., Zamudio, K.R., 2020. Coalescent- Smith, M.A., Green, D.M., 2005. Dispersal and the metapopulation paradigm in
based species delimitation is sensitive to geographic sampling and isolation by amphibian ecology and conservation: are all amphibian populations
distance. Syst. Biod. 18, 269–280. https://doi.org/10.1080/ metapopulations? Ecography 28, 110–128. https://doi.org/10.1111/j.0906-
14772000.2020.1730475. 7590.2005.04042.x.
Nascimento, J., Lima, J.D., Suárez, P., Baldo, D., Andrade, G.V., Pierson, T.W., Solís-Lemus, C., Knowles, L.L., Ané, C., 2015. Bayesian species delimitation combining
Fitzpatrick, B.M., Haddad, C.F.B., Recco-Pimentel, S.M., Lourenço, L.B., 2019. multiple genes and traits in a unified framework. Evolution 69, 492–507. https://
Extensive cryptic diversity within the Physalaemus cuvieri-Physalaemus ephippifer doi.org/10.1111/evo.12582.
species complex (Amphibia, Anura) revealed by cytogenetic, mitochondrial, and Stephens, M., Smith, N.J., Donnelly, P., 2001. A new statistical method for haplotype
genomic markers. Front. Genet. 10, 719. https://doi.org/10.3389/ reconstruction from population data. Am. J. Hum. Genet. 68, 978–989. https://doi.
fgene.2019.00719. org/10.1086/319501.
Nylinder, S., Lemey, P., De Bruyn, M., Suchard, M.A., Pfeil, B.E., Walsh, N., Talavera, G., Dincă, V., Vila, R., 2013. Factors affecting species delimitations with the
Anderberg, A.A., 2014. On the biogeography of Centipeda: a species-tree diffusion GMYC model: insights from a butterfly survey. Methods Ecol. Evol. 4, 1101–1110.
approach. Syst. Biol. 63, 178–191. https://doi.org/10.1093/sysbio/syt102. https://doi.org/10.1111/2041-210X.12107.
Ogilvie, H.A., Bouckaert, R.R., Drummond, A.J., 2017. StarBEAST2 brings faster species Thomassen, H., Ziade, C.F., 2020. Guia ilustrado de répteis e anfíbios da área de
tree inference and accurate estimates of substitution rates. Mol. Biol. Evol. 34, influência da Usina Hidrelétrica de Emborcação, first et. CEMIG, Belo Horizonte,
2101–2114. 10.1093/molbev/msx126. Brazil.
Olson, D.M., Dinersteins, E., Wikramanayake, E.D., Burgess, N.D., Powell, G.V.N., Vacher, J., Chave, J., Ficetola, F.G., Sommeria-Klein, G., Tao, S., Thébaud, C., Blanc, M.,
Underwood, E.C., D’Amico, J.A., Itoua, I., Strand, H.E., Morrison, J.C., Loucks, C.J., Camacho, A., Cassimiro, J., Colston, T.J., Dewynter, M., Ernst, R., Gaucher, P.,
Allnutt, T.F., Ricketts, T.H., Kura, Y., Lamoreux, J.F., Wettengel, W.W., Hedao, P., Gomes, J.O., Jairam, R., Kok, P.J.R., Lima, J.D., Martinez, Q., Marty, C., Noonan, B.
Kassem, K.R., 2001. Terrestrial ecoregions of the world: a new map of life on Earth. P., Nunes, P.M.S., Ouboter, P., Recoder, R., Rodrigues, M.T., Snyder, A., Marques-
Bioscience 51, 933–938. https://doi.org/10.1641/0006-3568(2001)051[0933: Souza, S., Fouquet, A., 2020. Large-scale DNA-based survey of frogs in Amazonia
TEOTWA]2.0.CO;2. suggests a vast underestimation of species richness and endemism. J. Biogeogr. 47,
Oswald, C.B., Lemes, P., Thomé, M.T.C., Pezzuti, T.L., Santos, F.R., Garcia, P.C.A., 1781–1791. https://doi.org/10.1111/jbi.13847.
Leite, F.S.F., Magalhães, R.F., 2022. Colonization rather than fragmentation explains Van Bocxlaer, I., Biju, S.D., Loader, S.P., Bossuyt, F., 2009. Toad radiation reveals into-
the geographical distribution and diversification of treefrogs endemic to Brazilian India dispersal as a source of endemism in the Western Ghats-Sri Lanka biodiversity
shield sky islands. J. Biogeogr. 49, 682–698. https://doi.org/10.1111/jbi.14320. hotspot. BMC Evol. Biol. 9, 131. https://doi.org/10.1186/1471-2148-9-131.
Otto-Bliesner, B.L., Marshall, S.J., Overpeck, J.T., Miller, G.H., Hu, A., CAPE Last Vasconcellos, M.M., Colli, G.R., Weber, J.N., Ortiz, E.M., Rodrigues, M.T., Cannatella, D.
Interglacial Project Members, 2006. Simulating artic climate warmth and icefield C., 2019. Isolation by instability: historical climate change shapes population
retreat in the last interglaciation. Science 311, 1751–1753. 10.1126/ structure and genomic divergence of treefrogs in the Neotropical Cerrado savanna.
science.1120808. Mol. Ecol. 28, 1748–1764. https://doi.org/10.1111/mec.15045.
Phillips, S.J., Anderson, R.P., Schapire, R.E., 2006. Maximum entropy modeling of Vaz-Silva, W., Maciel, N.M., Nomura, F., Morais, A.R., Batista, V.G., Santos, D.L.,
species geographic distributions. Ecol. Modell. 190, 231–259. 10.1016/j.ec Andrade, S.P., Oliveira, A.Â.B., Brandão, R.A., Bastos, R.P., 2020. Guia de
olmodel.2005.03.026. identificação das espécies de anfíbios (Anura e Gymnophiona) do estado de Goiás e
Pons, J., Barraclough, T.G., Gomez-Zurita, J., Cardoso, A., Duran, D.P., Hazell, S., Distrito Federal, Brasil Central, first ed. Sociedade Brasileira de Zoologia, Curitiba,
Kamoun, S., Sumlin, W.D., Vogler, A.P., 2006. Sequence-based species delimitation Paraná. 10.7476/9786587590011.
for the DNA taxonomy of undescribed insects. Syst. Biol. 55, 595–609. https://doi. Webb, S.D., 1991. Ecography and the Great American Interchange. Paleobiology 17,
org/10.1080/10635150600852011. 266–280. https://doi.org/10.1017/S0094837300010605.

17
R.F. Magalhães et al. Molecular Phylogenetics and Evolution 190 (2024) 107959

Wells, K. D., 2007. The ecology and behavior of amphibians, first ed. The University of Wynn, A., Heyer, W.R., 2001. Do geographically widespread species of tropical
Chicago Press, Chicago, Illinois. http://dx.doi.org/10.7208/chicago/ amphibians exist? An estimate of genetic relatedness within the neotropical frog
9780226893334.001.0001. Leptodactylus fuscus (Schneider 1799) (Anura Leptodactylidae). Trop. Zool. 14,
Werneck, F.P., Nogueira, C., Colli, G.R., Sites Jr., J.W., Costa, G.C., 2012. Climatic 255–285. https://doi.org/10.1080/03946975.2001.10531157.
stability in the Brazilian Cerrado: implications for biogeographical connections of Yeates, D.K., Seago, A., Nelson, L., Cameron, S.L., Joseph, L., Trueman, J.W.H., 2011.
South American savannas, species richness and conservation in a biodiversity Integrative taxonomy, or iterative taxonomy? Syst. Ent. 36, 209–217. https://doi.
hotspot. J. Biogeogr. 39, 1695–1706. https://doi.org/10.1111/j.1365- org/10.1111/j.1365-3113.2010.00558.x.
2699.2012.02715.x.

18

You might also like