You are on page 1of 14

| |

Received: 12 January 2021    Revised: 9 February 2022    Accepted: 16 February 2022

DOI: 10.1111/jbi.14370

RESEARCH ARTICLE

Biogeography and diversification of bare-­eyes, an endemic


Amazonian clade

Lincoln Carneiro1,2  | Tibério C. T. Burlamaqui1  | Alexandre Aleixo1,3  |


David C. Oren4 | José Maria Cardoso Da Silva5

1
Coord. de Zoologia. Museu Paraense
Emílio Goeldi, Pará, Brazil Abstract
Aim: Biotic interchange, speciation and extinction processes drive biotas assembling.
2
Faculdade de Educação do Campo,
Universidade Federal do Pará, Cametá,
Brazil
However, the evolutionary outcomes of those mechanisms are complex and difficult
3
Finnish Museum of Natural History, to discriminate. Here, we investigate how these processes affect avian diversification
University of Helsinki, Helsinki, Finland in tropical forest regions and test the relative roles of vicariant speciation and biotic
4
Ministério da Ciência, Tecnologia e
interchange on the assemblage of Amazonian biota through the reconstruction of the
Inovações do Brasil (MCTI), Brasilia, Brazil
5
Department of Geography and biogeographical history of bare-­eyes using molecular markers.
Sustainable Development, University of Location: Amazon.
Miami, Miami, USA
Taxon: Phlegopsis (Aves: Thamnophilidae).
Correspondence Methods: We carried out a phylogenetic reconstruction of Phlegopsis based on four
Lincoln Carneiro, Coordenação de
Zoologia, Museu Paraense Emílio Goeldi, mtDNA and nuclear markers of 52 individuals from 21 localities representing all rec-
Caixa Postal 399, CEP 66040-­170, Belém, ognized taxa in the genus. We estimated phylogenetic relationships using both gene
Pará, Brazil.
Email: lscarneiro@ufpa.br tree and species tree methods, demographic history, gene flow and divergence time
to reconstruct the biogeographical history of the genus.
Funding information
Fundação Amazônia Paraense de Amparo Results: Phylogenetic analyses recovered nine lineages delimited by some of the larg-
à Pesquisa; Marion E. Kenworthy-­Sarah H; est Amazonian rivers. Molecular dating and biogeographical studies showed that (1)
Swift Action Fund
most of the diversification in the genus occurred during the Quaternary; (2) vicariance
Handling Editor: Camila Ribas was the most critical biogeographical process driving the history of this group and (3)
two lineages are expanding their ranges, with one of them (P. n. nigromaculata) cross-
ing the boundaries of areas of endemism.
Main conclusions: Our results reinforce the notion that continental biotas are assem-
bled by alternating dispersal and vicariance events. In Phlegopsis, vicariance shaped
the distribution and differentiation of most lineages, but one event of post-­speciation
dispersal made two lineages sympatric in Western Amazonia. Dispersal events within
the Amazon are not random but are constrained by the characteristics of the spe-
cies that are expanding their ranges and the features of the places that receive such
species. Although most of the speciation events of Phlegopsis occurred during the
Quaternary, there is no reason to assume that such events were driven by climatic-­
vegetation changes associated with the Milankovitch cycles.

KEYWORDS
Amazon, dispersal, molecular systematics, neotropical region, quaternary, speciation

Journal of Biogeography. 2022;00:1–14. wileyonlinelibrary.com/journal/jbi © 2022 John Wiley & Sons Ltd.     1 |
|
2      CARNEIRO et al.

1  |  I NTRO D U C TI O N interspecific competition (Willis & Oniki, 1978). In fact, Phlegopsis are


considered as ‘professional’ or ‘obligate’ army ant followers (Willis &
In the Amazon, there are nine biogeographical regions separated Oniki, 1978) that get most of their food by regularly following army-­
by some of the longest and widest rivers in the region and take the ants (primarily Eciton burchellii and Labidus praedator) in the forest
shape of large peninsulas with their tips pointing to the Amazon understorey to capture arthropods that attempt to flee ant swarms
River. Each one of these biogeographical regions has a unique biota (Willis & Oniki, 1978; Willis & Oniki, 1992). Because they are relatively
characterized by many endemic species. Therefore, these regions heavy and aggressive passerines (c. 50 g), Phlegopsis take the central
can also be classified as areas of endemism (Borges & Silva, 2012; and best zones over the ant swarms, supplanting smaller bird species
Cracraft, 1985; da Silva et al., 2002). The species living in these areas or chasing them to peripheral zones (Willis, 1982).
of endemism have different ages, evolutionary histories and habitat In this paper, we test the hypothesis that speciation by vicariance
requirements and have been assembled over time by tracking the is more important than post-­speciation dispersal events to explain the
region's environmental changes (Bates, 2001; da Silva & Oren, 1996; current geographical distribution of Phlegopsis and thus contribute to
Haffer, 2008; Ribas et al., 2012). the understanding of the processes driving bird diversification in trop-
Biotas are assembled due to three main biogeographical pro- ical forest regions. To do so, we use molecular markers to assess (a) the
cesses: biotic interchange, speciation and extinction (Cracraft, 1994; phylogenetic relationships within the group, (b) the divergence ages of
Ricklefs & Schluter, 1993). Biotic interchange adds species to a region, these lineages and (c) the demographic history of these lineages.
while extinction reduces the number of species. In contrast, in situ
speciation adds endemic species to a region, distinguishing it from
other regions (da Silva et al., 2005). Although speciation has been con- 2  |  M ATE R I A L S A N D M E TH O DS
sidered as the primary factor shaping the composition of the biotas
of the Amazonian areas of endemism (Bates, 2001; Carneiro, Bravo, 2.1  |  Sampling and sequence analysis
et al., 2018; da Silva & Oren, 1996; Ribas et al., 2012), there is evi-
dence that biotic interchange between areas of endemism can also be We sampled 52 specimens representing all species and subspecies
important (Antonelli et al., 2018; Bates, 2001; da Silva & Oren, 1996). recognized in Phlegopsis (Figure 1, Appendix S1). These species and
Among vertebrate clades restricted to the Amazon, related subspecies are as follows: P. e. erythroptera, P. e. ustulata; P. borbae;
species usually replace each other across the region's areas of en- P. n. nigromaculata, P. n. bowmani, P. n. confinis and P. n. paraensis. We
demism, with no range overlap between them. This distribution pat- also included sequences of the five other Thamnophilidae genera
tern is generally interpreted as a result of the range fragmentation (Gymnopithys, Rhegmatorhina, Pithys, Phaenostictus and Willisornis) as
of a widespread ancestor due to vicariant events followed by the the outgroups, following Isler et al. (2014).
differentiation of the isolated populations into daughter species Total genomic DNA was extracted using the Sambrook and
(Cracraft,  1985; Haffer,  1969, 1974, 1987; Ribas et al.,  2012). In a Russell's  (1989) protocol from tissue and blood samples. We used
few groups, however, closely related species are sympatric. Because standard methods described elsewhere (Kocher et al.,  1989;
sympatric speciation is rare among birds (Phillimore et al.,  2008), Prychitko & Moore,  1997; Slade et al.,  1993) to amplify and ob-
these cases of sympatry are considered to be the result of post-­ tain sequences for three mitochondrial coding genes and a nuclear
speciation dispersal events (Chesser & Zink, 1994; Price et al., 2014; intron—­cytochrome b (Cyt B), cytochrome oxidase I (COI), NADH de-
Schweizer & Liu, 2018). Thus, studies on the transition between spa- hydrogenase subunit 2 (ND2) and beta-­fibrinogen intron 7 (BF7), for
tial segregation (allopatry) to coexistence (sympatry) within clades primers details, see Appendix S2, Table S2.1.
using phylogenetic approaches promise to reveal an underappreci- Electropherograms were inspected, assembled in contigs
ated mechanism of biotic assembly in Amazonia. and edited in Geneious 9.1.2 (http://www.genei​ous.com, Kearse
The genus Phlegopsis (bare-­eyes) is an excellent model to examine et al.,  2012). Sequences were aligned with MAFFT v. 6 (Katoh
the biogeographical processes behind the transition between allo- et al., 2002) using the default parameters and further inspected visu-
patry and sympatry within the Amazon. There are three reasons for ally. The software PHASE (Stephens & Donnelly, 2003) was used to
this. First, the lineages of this genus show both allopatry and sympatry. estimate the gametic phases of FIB7, assuming a threshold of poste-
In general, the seven named lineages (ranked as subspecies) replace rior probability of 0.7. Unphased sequences with probabilities below
one another across most Amazonian areas of endemism (Haffer, 1992; this threshold were also used in the analysis. Moreover, signs of past
Zimmer & Isler, 2020), but some lineages co-­occur in several localities recombination of the nuclear loci were tested using a PHI test (Bruen
in western Amazonia due to their ability of them to live in both terra et al., 2006) in Splits-­Tree4 (Huson & Bryant, 2006).
firme and seasonally flooded (várzea) forests (Willis, 1982; Willis, 1984;
Zimmer & Isler, 2020). Second, Phlegopsis are forest birds with limited
dispersal propensity. They have rounded wings and rarely cross open 2.2  |  Phylogeny estimation
spaces, making them a reliable indicator of environmental changes
in the Amazon forests (Willis & Oniki,  1978). Finally, Phlegopsis are We estimated phylogenetic relationships through Bayesian Inference
specialized birds, making lineages' co-­occurrence less likely due to (BI) using both MrBayes 3.2.1 (Ronquist & Huelsenbeck,  2003) and
CARNEIRO et al.       3|
A1
A2
70°W 60°W 50°W
A7
A6 P. n. paraensis
*/* A4
A5
A8 N
*/* A3
C1 H4, H5 0°
C4
0.92/77 C2 P. n. bowmani
C3
*/* B1
C5 H3 Negro B1
B2 A4, A5, A6,
*/* B3 P. n. confinis Japurá
B4
H1, H2 B2, B3 A7, A8
B5
*/90 E1 D1, D2, C4, C5
E2 P. n. nigromaculata A B4, B5
E3 Solimões D3, D4
D11 F1, F2,
0.59/67 D10 Xingu
A1, A2, A3
*/* D9 F3, F4, Tapájos
*/* D6 Tocantins
D8
D7 P. n. nigromaculata B
F5, F6, Madeira
D5 F7, F8 C1, C2, C3
D2
D1 D8, D9,
D4
D3 D10, D11
*/* J5
*/90 J4 J5
J3 P. borbae
J1
J2 G1, G2 D5, D6, D7
*/* G1 E2
P. e. ustulata B
*/* F7 G2 J1, J2, J3, J4
*/96 F1
F2 10°S
F8 E3 0 200 400 km
*/* F3 P. e. ustulata A E1
*/* F6
F5
F4
H4
H5
0.73/55 H2 P. e. erythroptera
H1 Atlantic
H3
*/* Oneillornis lunulatus Ocean
85/62 Oneillornis salvini
*/* */* Gymnopithys bicolor
*/* Gymnopithys rufigula
*/84 Gymnopithys leucaspis
*/* Rhegmatorhina melanosticta
*/* Rhegmatorhina hoffmannsi
Rhegmatorhina gymnops
Willisornis poecillinotus South
*/* Pithys albifrons America
90/76 Pithys castaneus
Phaenostictus mcleannani

F I G U R E 1  Sampling localities and molecular phylogeny of Phlegopsis based on mtDNA genes Cytb, COI, ND2 and a nuclear intron, bf7
(2865 bp). Bayesian and maximum likelihood analysis yield the same topology. Numbers on nodes are posterior probabilities/ML bootstrap.
*indicates maximum support values

Maximum Likelihood (ML) in RAxML 7.2.8 (Stamatakis, 2006). Three We used a Yule speciation prior and a relaxed uncorrelated lognor-
independent runs conducted BI analysis with 107 generations, each mal clock for all genes (Drummond et al., 2006). The molecular dat-
generation with four Markov chains, one cold and three heated, with ing used a substitution rate of ND2 of 2.5% per million years (Smith &
one sample every 1000 generations. Output parameters were visu- Klicka, 2010) and estimated rates for the remaining markers. We spec-
alized using Tracer 1.6 (Drummond & Rambaut, 2007) to assess sta- ified a lognormal distribution and a relatively wide logarithmic stan-
tionarity and convergence (Effective Sample Size –­ ESS values >200). dard deviation of 0.2 for each gene, thus encompassing alternative
We discarded trees obtained before the chain reached stationarity to substitution rates (e.g. Lerner et al., 2011). We did three independent
estimate the topology and posterior probability of clades. We gener- runs with 5 × 107 generations each, with parameters sampled every
ated 1000 bootstrap pseudo-­replicates to evaluate the support for 5000 generations. The ESS values and the number of discarded sam-
ML. Both analyses used the unphased concatenated databank for all ples were inferred with the help of TRACER 1.5 (Rambaut et al., 2014).
genes with the best partition scheme and model of molecular evolu-
tion inferred by PartitionFinder 1.1.1 (Lanfear et al., 2012). The optimal
partition scheme and best-­fit models used on the BI are described in 2.4  |  Biogeographical analysis
detail in the supplementary material (Table S2.2, Appendix S2). This
concatenated phylogeny was used as a guide tree for further analysis To estimate the biogeographical history of Phlegopsis, we used
(Species tree, BioGeoBEARS and population genetics). Essentially, sup- ‘BioGeoBEARS (Biogeography with Bayesian (and likelihood)
ported reciprocally monophyletic groups that were also geographically Evolutionary Analysis’ in R Scripts; Matzke, 2013; http://cran. r-­proje​
structured were assumed to be hypothesized evolutionary units. ct.org/web/packa​ges/BioGe​oBEAR​S/index.html). BioGeoBEARS
models probabilistic biogeographical scenarios onto a user-­defined
phylogeny. We ran six different analyses including the Dispersal–­
2.3  |  Species tree and molecular dating Extinction Cladogenesis Model (DEC), a likelihood version of the
Dispersal–­V icariance Analysis (‘DIVALIKE’), and a version of the
We used *BEAST (Heled & Drummond,  2010), part of the BEAST Bayesian inference of historical biogeography for discrete areas
2.5 package (Bouckaert et al., 2019), to reconstruct a time-­calibrated (BAYAREALIKE), as well as ‘+J’ versions of these three models,
species tree, using all individuals of Phlegopsis lineages recovered by which include founder event speciation (Matzke,  2013). In these
our previous analyses. The partitions and models used on the spe- analyses, we used our multilocus species tree to infer the ancestral
cies tree are described in detail in the supplementary material (Table range probabilities by defining the eight Amazonian areas of ende-
S2.2, Appendix S2). mism where Phlegopsis is found as biogeographical units: IM = Imeri;
|
4      CARNEIRO et al.

NP = Napo; JA = Jaú; IN = Inambari; RO = Rondônia; TA = Tapajos; likelihood-­ratio tests different models implemented in the IMA2 L
XI = Xingu; and BE = Belém (sensu Borges & Silva, 2012; da Silva mode.
et al.,  2002). Although we have no molecular samples of P. n. ni- To analyse population size dynamics through time for both mi-
gromaculata from Napo, phenotypic data (acoustic and morphologi- tochondrial and nuclear loci combined, we estimated extended
cal) suggest a substantial similarity between the populations of P. n. Bayesian skyline plots (EBSPs, Heled & Drummond, 2010) available
nigromaculata A living in Napo and Inambari (Carneiro et al., in prep). in the BEAST 2.5 package (Bouckaert et al., 2019). EBSP analyses in-
We included this information in our biogeographical setting. cluded lineages recovered by concatenated phylogeny. The best-­fit
We set the maximum number of areas to eight. For the node pie substitution model for each marker, substitution rates, clock rates,
charts of ancestral biogeographical ranges, we accepted probability priors and the MCMC run strategies was the same described for the
values in which the most likely area was >5% (distributions with <5% phylogenetic analyses.
probability of occurrence combined are the white portions in the
pie charts; Figure  2). Finally, we compared the six different mod-
els for statistical fit via comparison of Log-­Likelihood (ln L), Akaike 3  |  R E S U LT S
Information Criterion (AIC), ΔAIC and Akaike weight (ωi) values
(Wagenmakers & Farrell, 2004). 3.1  |  Sequence data and genetic diversity

We sequenced a total of 2861 bp (Table S2.3 in Appendix S2). The


2.5  |  Population genetics and nuclear locus showed no recombination (PHI test, p  =  0.534) and
historical demography higher intraspecific nucleotide diversity than that of the mtDNA
(Table S2.4 in Appendix S2). Phlegopsis lineages showed higher ge-
We used DnaSP 5.10.01 (Librado & Rozas, 2009) to perform Tajima's netic diversity for ND2 than other genes (HD and π; Table S2.4 in
D (Tajima, 1989), Fu's FS (Fu, 1997) and the Ramos-­Onsins and Rozas' Appendix S2).
R2 tests (Ramos-­Onsins & Rozas,  2002) under 1000 coalescent
simulations. We estimated the genealogical relationships among
haplotypes using median-­joining networks (Bandelt et al.,  1999) 3.2  |  Phylogenetic analyses
as implemented in Haploviewer (http://www.cibiv.at/~greg/haplo​
viewer). For haplotype network, we removed the sequences and The concatenated tree supported that (1) Phlegopsis has nine well-­
parts of the fragment containing missing data. We used a Bayesian defined lineages, (2) these lineages are reciprocally monophyletic
Analysis of Population Structure (BAPS v 6.0.; Corander et al., 2008) and (3) Phlegopsis are a sister group to the clade formed by the
to estimate the level of geographical structure within P. nigromacu- genera Gymnopithys, Rhegmatorhina, Pithys, Phaenostictus and
lata and P. erythroptera separately. The probabilities of genetic clus- Willisornis (Figure  1; Harvey et al.,  2020; Isler et al.,  2014). The
ters (from K = 1 to K = 10) for each group were estimated using the bare-­eyes can be grouped into two well-­supported clades: the first
mtDNA dataset only and mixture and admixture analyses. The pro- one composed of P. e. erythroptera and P. e. ustulata (A and B line-
gram was run 10 times for each value of K. The admixture analysis ages), and a second one comprising P. borbae, P. n. nigromaculata
was performed using 1000 iterations, a minimum of one individual (A and B lineages), P. n. bowmani, P. n. confinis and P. n. paraensis
per population, 200 reference individuals for each population and (Figure 1). The first group is found in the western Amazon, with P. e
20 iterations of reference individuals. erythroptera occupying three areas of endemism (Imeri, Napo and
We used the isolation–­migration model (Hey & Nielsen, 2004; Jaú) and P. e ustulata (A and B) occupying one (Inambari) (Figure 1).
Nielsen & Wakeley, 2001) implemented in IMA2 (Hey, 2010) to esti- In the second group, P. borbae is restricted to one area of end-
mate levels of gene flow between neighbouring populations. IMA2 emism (Rondônia), P. n. nigromaculata A is found in Inambari and
estimates were obtained only for reciprocally monophyletic pop- Napo, P. n. nigromaculata B is restricted to Rondônia. P. n. bowmani
ulation pairs in adjacent areas. IMA2 analyses employed the HKY is endemic to Tapajós, P. n. confinis is endemic to Xingu, and, fi-
model (Hasegawa et al., 1985) for all markers; an inheritance scale nally, P. n. paraensis is restricted to Belém.
of 0.25 for the mtDNA and 1 for nDNA; a substitution rate for the
ND2 gene of 2.5% sequence divergence per million years per gen-
eration (estimating those of cytb, COI and BF7) and a 1.4 year gen- 3.3  |  Species tree and divergence time estimates
eration time (Bird et al., 2020). We did several runs to establish the
best priors for effective population sizes, time of divergences and The first divergence within Phlegopsis was between the ancestor
migration parameters. Then, we performed three final runs with 105 of P. erythroptera and P. ustulata and the ancestor of all other line-
5 6
generations as burn-­in, 10 trees sampled during 10 generations ages around 1.6 Ma (time divergence estimate ranging from 0.75 Ma
and 20 chains. Finally, to test whether a model of isolation without to 3.02 Ma; Figure 2 and Figure S3.1). The next split separated the
gene flow fit the data better than a model with gene flow, we used ancestor of P. borbae from the ancestor of all subspecies of P. ni-
Nielsen and Wakeley's approach (Nielsen & Wakeley, 2001) and the gromaculata clade around 1.2  Ma (0.51  Ma–­1.95  Ma). Then, P. n.
CARNEIRO et al. |
      5

ustulata diverged from P. n. erythroptera (0.8 Ma, 0.39 Ma–­1.04 Ma). possible combinations. However, these results are more realistic and
Finally, P. n. nigromaculata diverged from the ancestor of P. n. bowm- free from previous constraints that could bias biogeographical anal-
ani–­P. n. confinis–­P. n. paraensis around 0.6  Ma (0.43–­0.86  Ma), P. ysis. The DIVALIKE model suggests a scenario in which the lineages
n. bowmani split from the ancestor of P. n. confinis–­P. n. paraensis at of Phlegopsis result from genetic divergence due to the successive
0.45 Ma (0.34–­0.66 Ma), and P. n. confinis and P. n. paraensis diverged subdivision of the populations of an immediate common ancestor
at 0.4 Ma (0.19–­0.57 Ma). by vicariance with a few cases of sympatry between lineages due to
post-­speciation dispersal events (Figure 2).

3.4  |  Biogeographical analysis


3.5  |  Population genetics and
Of the six biogeographical models evaluated, the best fit to our historical demography
dataset pertained to +J models (Appendix S3). All the +J models
show similar values (Table 1). Nevertheless, the +J versions of the The mtDNA or nuclear markers found no shared haplotypes among
different models generated by BioGeoBEARS are doubtfully ap- the nine Phlegopsis lineages (Figure 3a). Gene flow analyses contrast-
plied to non-­island groups and can lead to unrealistic results (Ree ing only Phlegopsis lineages potentially in contact across riverine bar-
& Sanmartín, 2018). Thus, we chose the DIVALIKE model, which is riers detected migration rates not statistically different from zero
the best fit for our data without the ‘+J' component (Figure 2). The (Table S2.5 in Appendix S2), indicating strong genetic isolation.
pie charts have multiple uncertainties that reflect the complexity of The P. n. nigromaculata clusters detected by BAPS were not the
the group's biogeographical pattern and the way we implemented same recovered by phylogenetic analyses. BAPS separated P. n. ni-
the analysis by allowing the software to search the probabilities for gromaculata into three lineages and not five, as the concatenated
all eight defined areas without restrictions. This approach gener- tree. The BAPS analysis recovered an admixture event involving P. n.
ates less conclusive results mainly in the old nodes due to the many nigromaculata A and P. n. bowmani (Figure 3b), as the migrations tests

NPIMJA P. e. erythroptera
IMJAIN
NPIMIN
NPIMJAIN
BioGeoBEARS IN P. e. ustulata A
Model: DIVALIKE
8 areas max.
Atlantic d (dispersal)=0.0713; IN P. e. ustulata B
e (extinction)=0;
Ocean

j (founder)=0;
LnL=−23.98
RO P. borbae

NPINROTAXIBE
ROTAXIBE INRO NPIN P. n. nigromaculata A
RO NPINRO NPRO

RO P. n. nigromaculata B
INROTAXIBE
Imeri NPINROTAXIBE NPROTAXIBE
Napo
Jaú Belém
TA P. n. bowmani

Inambari
Tapajós
Xingu TAXIBE XI P. n. confinis
Rondônia
XIBE
BE P. n. paraensis

1.5 1.0 0.5 0 Mya

F I G U R E 2  Biogeographical history for the Phlegopsis antbirds generated by BioGeoBEARS. The best-­fit model without J for our
reconstruction was DIVALIKE (ln L = −23.97, AIC = 51.96; see text for details). Colours represent the current distribution of taxa and
subclades. Node pie charts represent the likelihoods of ancestral states. Uncertainty in ancestral range reconstruction is shown where
distributions with <5% probability of occurrence are combined into white portions in pie charts
|
6      CARNEIRO et al.

TA B L E 1  Models, parameters and scores from each of the analyses in BioGeoBEARS. Dispersal (d), extinction (e), founder (J), values of
log-­likelihood (ln L), Akaike information criterion (AIC), ΔAIC value and Akaike weight (ωi)

Rank Model Parameters d e j Ln L AIC ΔAIC ωi

1 DIVALIKE+J 3 0.0259 1.00e-­12 0.212 −21.16 48.33 0 5.57e-­01


2 BAYAREALIKE+J 3 0. 0165 1.00e-­07 0.318 −22.24 50.49 2.16 1.89e-­01
3 DEC + J 3 0.0215 1.00e-­12 0.367 −22.39 50.79 2.46 1.62e-­01
4 DIVALIKE 2 0.0712 1.00e-­12 0 −23.97 51.96 3.63 9.07e-­02
5 DEC 2 0.1155 3.44e-­01 0 −29.74 63.49 15.16 2.84e-­0 4
6 BAYAREALIKE 2 0.2238 1.8372 0 −31.93 67.87 19.54 3.18e-­05

(a)

(b)

F I G U R E 3  (a) Median-­joining haplotype networks estimated for each sequenced gene. Cyt b, 52 individuals, 453 bp; ND2, 49 individuals,
906 bp; COI, 52 individuals, 559 bp; and bf7, 52 individuals, 853 bp; (b) population clusters defined by the Bayesian analysis of population
structure (BAPS) are represented by thick black lines. Colours represent the populations recovered by phylogenetic analyses (see text for
details). Each bar corresponds to an individual (code and colours follow Figure 1)

were not statistically significant. This result may indicate retention The EBSP analyses indicated increasing population sizes for
of ancestral polymorphism, but a contact at the southern edge of most lineages, except P. erythroptera (Figure  4). There is evidence
the Amazon cannot be ruled out, as recovered by Weir et al. (2015) of gradual population growth in the last 50,000–­500,000 years,
and Cronemberger et al. (2020) and Dantas et al. (2021). Regarding with P. n. nigromaculata B and P. borbae exhibiting the sharpest in-
P. erythroptera, the results agreed with the phylogenetic analyses, crease among all populations (Figure  4). P. erythroptera showed no
except for an additional separation within P. erythroptera (individual significant deviations from the null hypothesis of neutrality and con-
H5, Figure 3b). stant population size for any marker, while P. ustulata had significant
CARNEIRO et al.

Phlegopsis nigromaculata paraensis Phlegopsis nigromaculata confinis Phlegopsis nigromaculata bowmani Phlegopsis nigromaculata A

3.5

3.0
4

3.0

2.5
2.5
3

2.0
2.0
2

1.5
2

1.5

1.0

Population size scalar


Population size scalar
Population size scalar
Population size scalar

1.0
1

0.5
0.5
0
0

0.0
0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.00 0.05 0.10 0.15 0.20 0.25 0.0 0.1 0.2 0.3 0.4

Time (M.y.) Time (M.y.) Time (M.y.) Time (M.y.)

Phlegopsis nigromaculata B Phlegopsis erythroptera erythroptera Phlegopsis borbae


Phlegopsis erythroptera ustulata A

6
4

2.5

5
4
3

2.0

4
3

3
1.5
2

2
1.0

Population size scalar


Population size scalar
Population size scalar

Population size scalar


1

1
0.5

0
0

0.0

0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.5 1.0 1.5 2.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Time (M.y.) Time (M.y.) Time (M.y.)
Time (M.y.)

F I G U R E 4  Historical demographic inferred through extended Bayesian skyline plots based on mtDNA and nuclear sequence data. Dashed lines represent median values, while the solid lines
delimiting the grey area correspond to 95% confidence intervals. The X-­axis corresponds to time in million years before present. The Y-­axis represents Nₑτ in a logarithmic scale
|
      7
|
8      CARNEIRO et al.

deviations only for ND2. P. borbae and all lineages classified under P. haplotype networks, indicated that diversification of crown bare-­
nigromaculata, except P. nigromaculata A, showed expanding popula- eyes is relatively recent, dating to less than ~2.0 million years ago
tions for at least one locus (Table S2.4 in Appendix S2). However, our (Ma). However, the dates retrieved by us in the current study are
demographic and population results suggest a trend and should be quite different from those recovered by previous studies for
analysed with caution due to the relatively low number of localities Phlegopsis (Harvey et al., 2020; Silva et al., 2019). The dataset used
sampled from each population. here is virtually the same used in Silva et al. (2019) and the difference
between the two estimates is undoubtedly related to the choice of
the markers used in the calibration of the molecular clock (less than
4  |  D I S C U S S I O N half of the Cytb in Silva et al. (2019) against the complete ND2 in the
present study). Our option here was to minimize potential problems
We used genetic analyses at both phylogeographical and phyloge- with missing data, which are known to impact the time estimation
netic levels to reconstruct the biogeographical history of Phlegopsis. (Zheng & Wiens, 2015).
Our results showed that (1) Phlegopsis comprises nine well-­marked Regarding the date estimation by Harvey et al.  (2020) for the
lineages associated with Amazonian areas of endemism; (2) most clade, the differences are likely related to the phylogenetic scale
of the diversification in this genus occurred during the Quaternary; of the two studies and to the different taxa included. The study by
(3) vicariance is the most critical biogeographical mechanism driv- Harvey et al. (2020) included only one individual of each recognized
ing the history of this group; (4) two lineages are expanding their species and, as a consequence, did not consider large genetic vari-
ranges, with one of them (P. n. nigromaculata A) crossing the bound- ation that exists within each taxon. In general, the ages retrieved
aries of areas of endemism. Altogether, these four findings shed in our estimate for Phlegopsis are similar to those found by Ribas
light on some important and hitherto underappreciated aspects of et al.  (2018), for Rhegmatorhina, a phylogenetic and ecologically
Amazonian biogeography. close group to Phlegopsis, and to most of the dates proposed for pas-
serines in the Amazon basin (Silva et al., 2019).
The haplotype networks showed a structure in mitochondrial
4.1  |  Phlegopsis has nine well-­marked lineages DNA but not in the nuclear marker, consistent with their relative co-
associated with areas of endemism alescent times (Hung et al., 2017; Zink & Barrowclough, 2008). Thus,
our results are consistent with those found in other Amazonian
We found that Phlegopsis has at least nine genetically well-­marked clades (Hylopezus, Myrmothera, Psophia and Rhegmatorhina), in
lineages with no evidence of gene flow between them, with two ex- which a nuclear marker was only capable of resolving the oldest
ceptions (P. erythroptera and P. n. nigromaculata A; Figure 2), each is divergences (Carneiro, Bravo, et al., 2018; Ribas et al., 2012; Ribas
restricted to one area of endemism. This general pattern is similar to et al., 2018). The nuclear marker also indicated a recent divergence
several other forest groups endemic to the region (da Silva et al., 2002; among borbae and the ancestor of nigromaculata-­bowmani-­confinis-­
da Silva & Oren,  1996; Fernandes et al.,  2014; Ribas et al.,  2012; paraensis compared to the distance separating the ancestor of this
Santana et al.,  2020) and reinforces the hypothesis advanced by group from the ancestor of erythroptera-­ustulata. Phlegopsis lineages
Wallace (1852) that the areas of endemism delimited by some of the are common monogamous understorey birds and are expected to
longest and widest of the region's rivers are a valid and useful frame- have a large effective population sizes (Ne), relative to bigger spe-
work upon which biogeographical hypotheses can be proposed and cies, thus further suggesting that nuclear loci would not be pre-
tested (Bates, 2001; Cracraft, 1985; da Silva et al., 2002; da Silva & dicted to coalesce in Phlegopsis over this short time-­period (Ribas
Garda, 2010; da Silva & Oren, 1996; Ribas et al., 2012). However, we et al., 2018; Willis & Oniki, 1978). Thus, estimating age and popula-
found that two individuals of P. erythroptera (H4 and H5) collected in tion size using mitochondrial DNA is more appropriate than a single
the same locality on the east bank of the middle-­upper Negro River or a few nuclear markers for recovering evolutionary relationships
showed marked differences in mtDNA. This piece of information indi- within species complex at this taxonomic scale (Ribas et al., 2018).
cates the possibility of gene flow across the river, an unexpected find- Migration after divergence was not tested and if it occurred, it may
ing for a rainforest understorey antbird (Burney & Brumfield, 2009; have affected the time estimates produced by the species tree anal-
Claramunt et al., 2012; Smith et al., 2014). The gene flow in erythrop- ysis which would tend to be more recent. The *BEAST analysis does
tera (H4 and H5) may be related to changes in the main channel of the not consider the possibility of migration after isolation, considering
Negro River (Cremon et al., 2016). all similarities among species as ancestral polymorphism (Heled &
Drummond, 2010).
Quaternary speciation in the Amazon is usually interpreted as
4.2  |  The origin and diversification of Phlegopsis an outcome of the large-­scale climatic-­vegetational changes caused
occurred during the Quaternary by the Milankovitch cycles (Haffer, 1987; Haffer, 1993). That is the
foundation of the refuge model (Haffer, 1969), which has been the
Despite their genetic distinctness and lack of gene flow, the small dominant paradigm to explain avian diversification in the region in
genetic distances between lineages, even within the mitochondrial the last decades (Silva et al.,  2019). However, the original refuge
CARNEIRO et al. |
      9

model is based on the assumption that the region's forests were re- and P. n. confinis-­P. n. paraensis is linked to the Xingu, and the sep-
placed by open vegetations across the region during glacial periods aration between P. n. confinis and P. n. paraensis with the Tocantins.
(Rocha & Kaefer,  2019; da Silva & Garda,  2010). According to the In addition to vicariant events between areas of endemism, we also
model, speciation events, particularly those that occurred during the identified a split (separating P. borbae from P. n. nigromaculata–­P. n.
Quaternary, would be directly related to climate and vegetational bowmani–­P. n. confinis–­P. n. paraensis) within the area of endemism
changes caused by glacial cycles (Arruda et al., 2018; Haffer, 1969). Rondônia. P. borbae and P. n. nigromaculata B are found together in a
However, several studies have documented an intense geological few sites with the area of endemism Rondônia, but the extension of
dynamic in the region during the Quaternary as a consequence of this sympatry zone needs to be determined. In general, our results
neotectonics (Latrubesse & Rancy,  2000; Rossetti,  2014), changes indicate that the vicariant events that shaped the evolutionary his-
in sea level (Hoorn et al.,  2016; Irion & Kalliola,  2010) and fluvial tory of Phlegopsis started along the Madeira drainage and advanced
dynamics (Räsänen et al.,  1987; Rossetti,  2014). Furthermore, pa- progressively towards west and east. This west–­east pattern was
laeoriver dynamics that do not correspond to current river config- also identified for other groups of terrestrial vertebrates such as pri-
urations could also be related to diversification events in Amazon mates (Carneiro, Sampaio, et al., 2018; Lynch-­Alfaro et al., 2014), but
basin if they acted as semipermeable dispersal barriers through time has been especially reported for birds (Azuaje-­Rodríguez et al., 2020;
(Fernandes et al., 2014; Silva et al., 2019; Weir et al., 2015). Finally, Ferreira et al.,  2017; Ribas et al.,  2012; Santana et al.,  2020; Silva
all palaeoecological data assembled to date have failed to show a et al., 2019).
large-­scale replacement of forests by open vegetation across the The vicariant events within Rondônia are consistent with
entire region during the glacial periods (e.g. Irion & Kalliola,  2010; the patterns found in other groups of vertebrates, such as birds
Kastner & Goni, 2003). In fact, such studies show only limited vege- (Fernandes et al.,  2014; Ferreira et al.,  2017; Ribas et al.,  2018;
tation changes along the transition between the Amazon and other Santana et al.,  2020; Willis,  1969) and tamarins of the genus Mico
ecological regions as well as around the existing Amazonian savanna (Roosmalen et al.,  2000), which show several speciation events
patches (Irion & Kalliola,  2010). In conclusion, there is no palaeo- within this area of endemism. These cases suggest that Rondônia is
ecological evidence to assume that Quaternary speciation events in a composite area of endemism, with a complex and dynamic biolog-
the Amazon are necessarily a consequence of vegetation changes ical history that requires further investigation (Bates, 2001; da Silva
caused by global climate changes. et al., 2005; da Silva & Oren, 1996; Ferreira et al., 2017). This area
of endemism is located in the contact zone between the old terrains
associated with the Pre-­C ambrian Brazilian plateau and the most re-
4.3  |  Vicariance is the most important cent Tertiary and Quaternary terrains of western Amazon (da Silva
biogeographical process factor driving the et al., 2005), suggesting that speciation events within the region are
history of Phlegopsis driven by geological events (Silva et al., 2019). Recent studies show-
ing a particular geological dynamic and several drainage anomalies
Our results showed a complex biogeographical history for the lin- within this area of endemism support this hypothesis (Hayakawa &
eages of Phlegopsis that includes both vicariance and dispersal Rossetti, 2015; Latrubesse, 2002).
processes, but that vicariance was the dominant one. Unlike other
biogeographical studies in Amazon basin (e.g. Carneiro, Bravo,
et al., 2018; Ferreira et al., 2017; Silva et al., 2019), we did not re- 4.4  |  One lineage (nigromaculata) expanded its
strict the maximum number of areas because there is no evidence range to other areas of endemism
to safely limit the distributions of ancestral nodes to a specific area
of endemism or group of areas of endemism. In fact, it is perfectly Although vicariance was the dominant process shaping the cur-
plausible that the ancestor of Phlegopsis was distributed throughout rent distribution of the Phlegopsis lineages, our results support the
most of the current ranges of its descendant species. Furthermore, hypothesis that some lineages are expanding their ranges. EBSP
maintaining the maximum number of areas does not restrict models analyses suggest that P. n. bowmani and P. n. nigromaculata A and B
with fewer areas, but limiting the maximum number could prevent have expanded their ranges while the other lineages remained sta-
the search for vicariant models that include more areas. ble (Figure 4). P. n. bowmani is apparently expanding its range within
Vicariant events within Phlegopsis occurred mainly along the one area of endemism. In contrast, P. n. nigromaculata may have
boundaries of the eight areas of endemism, indicating that such dispersed first within Rondônia and became sympatric with borbae
events can be associated with the largest and widest rivers of the (Willis, 1982). Then, it may have crossed the Madeira and colonized
region. The vicariant event that separated the P. ustulata–­P. eryth- most Inambari and Napo becoming sympatric with P. e. ustulata.
roptera clade from all other lineages is associated with the Madeira, There are three general dispersal models to explain the range ex-
while the divergence between P. ustulata and P. erythroptera is associ- pansion of one lineage (review in Morrone, 2009): (1) jump disper-
ated with the upper Amazon/Solimões River. In eastern Amazon, the sal, which are random movements of organisms across barriers that
split between P. n. nigromaculata and P. n. bowmani–­P. n. confinis–­P. allow the establishment of species in very distant areas, (2) diffusion,
n. paraensis is associated with the Tapajós, separating P. n. bowmani defined as a gradual and fast range expansion through movements
|
10      CARNEIRO et al.

of populations crossing adjacent habitats over several generations dispersal ability across large rivers (Crouch et al., 2019). Therefore, P. n.
without evolutionary changes, and (3) secular migration, which is nigromaculata expanded its range across western Amazon by crossing
similar to diffusion, but slower and with evolutionary changes. Of the large rivers of the region via ‘passive dispersal’ associated with the
these three models, jump dispersal is unlikely to P. n. nigromaculata fluvial dynamics of the white-­water rivers.
because of its ecological traits. Between diffusion and secular mi-
gration models, our results support the latter. The main reason is
that populations of P. n. nigromaculata in western Amazon (lineage 4.5  |  Lessons for Amazonian biogeography
A) show a slight but consistent genetic difference in relation to its
putative source population in Rondônia (lineage B). In general, our results reinforce the notion that continental biotas are
Despite its limited dispersal potential, our results show that some assembled by dispersal events, followed by vicariance events, fol-
P. n. nigromaculata populations expanded their ranges in the last lowed by new dispersal events (Cracraft, 1994). Because the diversifi-
150,000–­200,000 years (Figure 4). What were the underlying mecha- cation of Phlegopsis is recent, we were able to reveal some elements of
nisms that could enable such expansion? Based on the current knowl- its biogeographical history. Similar to other studies (e.g. Bates, 2001;
edge in the ecology and behaviour of the lineages of Phlegopsis, we da Silva & Oren, 1996; Ribas et al., 2012), we found that vicariance
suggest that the range expansion of P. n. nigromaculata across western was the most important event shaping the distribution and differ-
Amazon was enabled by two mechanisms: (1) the ability of P. n. nigro- entiation within Phlegopsis. However, we have also identified that
maculata to live in seasonally flooded várzea forests and (2) passive secular migration can cause sympatry between closely related spe-
dispersal across rivers due to fluvial dynamics (Bicudo et al.,  2019; cies and thus increase the species diversity of areas of endemism over
Pupim et al., 2019; Rossetti, 2014; Sousa-­Neves et al., 2013). time (da Silva et al., 2005; D'Horta et al., 2013; Soares et al., 2019;
The ability to live in várzea forests is a recent adaptation within Santana et al., 2020). Such dispersal events are not random, but they
bare-­eye antbirds because it is restricted to the clade P. n. nigromac- are constrained by both the characteristics of the lineage that is ex-
ulata–­P. n. bowmani–­P. n. confinis–­P. n. paraensis. Across most of their panding its range as well as the characteristics of the places upon
ranges, lineages of this clade inhabit both terra firme and season- which the lineage is expanding (Cracraft, 1994; Crouch et al., 2019). In
ally flooded (várzea) forests (Willis,  1984). However, in the western the case of Phlegopsis, the successful expansion of P. n. nigromaculata
Amazon, P. n. nigromaculata is found only in várzea forests and the tran- westwards is a consequence of its ability to survive in várzea forests
sition of this forest to terra firme forests (Zimmer & Isler, 2020). In con- as well as the influence of the region's intense fluvial dynamics that
trast, P. e. ustulata and P. e. erythroptera are associated mostly with terra enables ‘passive’ dispersal across large rivers. Although most of the
firme forests (Zimmer & Isler, 2020). Thus, P. n. nigromaculata is rarely diversification of Phlegopsis occurred during the Quaternary, there
found in the same habitat as P. e. ustulata and P. e. erythroptera, even is no reason to assume that such processes were driven by climatic-­
though their ranges overlap. In the few localities where P. e. ustulata or vegetation changes associated with the Milankovitch cycles such as
P. e. erythroptera meet with P. n. nigromaculata over army-­ant swarms, proposed by the refuge model (Haffer,  1993). Because several of
P. n. nigromaculata occupies a subordinate position (Willis, 1982). Thus, the patterns that we described are also shared with other clades,
all evidence available suggests that P. n. nigromaculata expanded its we believe that, complex palaeoecological history notwithstanding,
range across western Amazon by occupying a habitat that had never the biogeographical patterns of present-­day organisms can be organ-
been used by the lineages of Phlegopsis endemic to that large region. ized out around a few general patterns that reveal the most general
This ecological plasticity could explain the fact that P. n. nigromaculata chapters of the history of Amazonian biota during the Quaternary
lineages, particularly in the east, show a distribution pattern different (Bates, 2001; da Silva & Oren, 1996; Ribas et al., 2012).
from that recovered for terra firme species (Ribas et al., 2012), and also
different from the patterns that have recently been recovered for sea- AC K N OW L E D G E M E N T S
sonally flooded forests species (Choueri et al., 2017; Thom et al., 2020). We thank the curator and curatorial assistants of the Museu
Várzea forests are associated with white-­water rivers, which re- Paraense Emílio Goeldi collection for allowing us to sequence sam-
ceive their names because they transport large quantities of sediments ples under their care. Specimens sequenced in this study were ob-
generated in the Andes (Sioli, 1984). These rivers are very dynamic and tained along several decades through collecting permits issued by
exhibit frequent lateral migration that transfers lands from one side of national environmental agency of Brazil. L.S.C. was supported by
the river to the other (Bicudo et al., 2019; Junk et al., 2011; Räsänen FAPESPA fellowship during the study (FAPESPA grant #021/2016).
et al., 1987). Because it is close to the Andes, western Amazon has the J. M. C. S. is supported by the University of Miami and the Swift
largest density of white-­water rivers than any other Amazonian re- Action Fund. New collection permits from the Conselho Nacional de
gion (Junk et al., 2011). In addition, recent studies have demonstrated Desenvolvimento Tecnológico e Científico—­CNPq and permission to
that because rivers in western Amazon have high sediment loads, they collect in protected areas from the Instituto Chico Mendes—­ICMBio
experience annual lateral migration rates that are higher than those are also gratefully acknowledged.
of rivers with lower sediment loads in other Amazonian drainages
(Constantine et al.,  2014). This geomorphological process, in turn, C O N FL I C T O F I N T E R E S T
can lead to the ‘passive dispersal’ of understorey birds with limited The authors declare no conflict of interest.
CARNEIRO et al. |
      11

DATA AVA I L A B I L I T Y S TAT E M E N T Carneiro, J., Sampaio, I., Silva-­júnior, J. D. S., Farias, I., Hrbek, T., Pissinatti,
All DNA sequences generated for this study are available from A., Silva, R., Martins-­junior, A., Boubli, J., Ferrari, S. F., & Schneider,
H. (2018). Phylogeny, molecular dating and zoogeographic his-
GenBank: https://www.ncbi.nlm.nih.gov/genbank/. Additional in-
tory of the titi monkeys (Callicebus, Pitheciidae ) of eastern Brazil.
formation and datasets, including newick trees and scripts used in Molecular Phylogenetics and Evolution, 124, 10–­15. https://doi.
ancestral range estimation are available from Dryad: https://doi. org/10.1016/j.ympev.2018.03.001
org/10.5061/dryad.3bk3j9kmw. Carneiro, L., Bravo, G. A., Aristizábal, N., Cuervo, A. M., & Aleixo, A.
(2018). Molecular systematics and biogeography of lowland
antpittas (Aves, Grallariidae): The role of vicariance and disper-
ORCID sal in the diversification of a widespread neotropical lineage.
Lincoln Carneiro  https://orcid.org/0000-0002-7680-2291 Molecular Phylogenetics and Evolution, 120, 375–­389. https://doi.
Tibério C. T. Burlamaqui  https://orcid.org/0000-0002-0264-5780 org/10.1016/j.ympev.2017.11.019
Chesser, R. T., & Zink, R. M. (1994). Modes of speciation in birds -­ a
Alexandre Aleixo  https://orcid.org/0000-0002-7816-9725
test of Lynch'S method. Evolution, 48(2), 490–­497. https://doi.
José Maria Cardoso Da Silva  https://orcid.org/0000-0002-7229-6694 org/10.2307/2410107
Choueri, É. L., Gubili, C., Borges, S. H., Thom, G., Sawakuchi, A. O.,
REFERENCES Soares, E. A. A., & Ribas, C. C. (2017). Phylogeography and pop-
ulation dynamics of antbirds (Thamnophilidae) from Amazonian
Antonelli, A., Zizka, A., Carvalho, F. A., Scharn, R., Bacon, C. D., Silvestro,
fluvial islands. Journal of Biogeography, 44, 2284–­2294. https://doi.
D., & Condamine, F. L. (2018). Amazonia is the primary source of neo-
org/10.1111/jbi.13042
tropical biodiversity. Proceedings of the National Academy of Sciences,
115(23), 6034–­6039. https://doi.org/10.1073/pnas.17138​19115 Claramunt, S., Derryberry, E. P., Remsen, J. V., Jr., & Brumfield, R. T.
Arruda, D. M., Schaefer, C. E. G. R., Fonseca, R. S., Solar, R. R. C., & (2012). High dispersal ability inhibits speciation in a continen-
Fernandes-­Filho, E. I. (2018). Vegetation cover of Brazil in the last tal radiation of passerine birds. Proceedings of the Royal Society B:
21 ka: New insights into the Amazonian refugia and Pleistocenic arc Biological Sciences, 279, 1567–­1574.
hypotheses. Global Ecology and Biogeography, 27(1), 47–­56. https:// Constantine, J. A., Dunne, T., Ahmed, J., Legleiter, C., & Lazarus, E. D.
doi.org/10.1111/geb.12646 (2014). Sediment supply as a driver of river meandering and flood-
Azuaje-­Rodríguez, R. A., Weckstein, J. D., Dispoto, J. H., Patel, S., plain evolution in the Amazon Basin. Nature Geoscience, 7(12), 899–­
Cacioppo, J. A., Bates, J. M., Silva, S. M., & Aleixo, A. (2020). Molecular 903. https://doi.org/10.1038/ngeo2282
systematics of the Amazonian endemic genus Hylexetastes (Aves: Corander, J., Marttinen, P., Sirén, J., & Tang, J. (2008). Enhanced
Dendrocolaptidae): Taxonomic and conservation implications. Ibis, Bayesian Modelling in BAPS Software for Learning Genetic
162(1), 119–­136. https://doi.org/10.1111/ibi.12693 Structures of Populations. BMC Bioinformatics, 9, 539. https://doi.
Bandelt, H.-­J., Forster, P., & Röhl, A. (1999). Median-­joining networks for org/10.1186/1471-­2105-­9-­539.
inferring intraspecific phylogenies. Molecular Biology and Evolution, Cracraft, J. (1985). Historical biogeography and patterns of differenti-
16(1), 37–­48. https://doi.org/10.1146/annur​ev.es.18.110187.002421 ation within the south American avifauna: Areas of endemism.
Bates, J. M. (2001). Avian diversification in Amazonia: Evidence for his- Ornithological Monographs, 36, 49–­8 4.
torical complexity and a vicariance model for a basic pattern of di- Cracraft, J. (1994). Species diversity, biogeography, and the evolution of
versification. In I. Viera, M. A. D'incão, J. M. C. Silva, & D. Oren biotas. Integrative and Comparative Biology, 34(1), 33–­47. https://
(Eds.), Diversidade Biológica e Cultural da Amazônia (pp. 119–­137). doi.org/10.1093/icb/34.1.33
Museu Paraense Emílio Goeldi. Cremon, É. H., Rossetti, D. D. F., Sawakuchi, A. D. O., & Cohen, M. C. L.
Bicudo, T. C., Sacek, V., de Almeida, R. P., Bates, J. M., & Ribas, C. C. (2016). The role of tectonics and climate in the late quaternary evo-
(2019). Andean tectonics and mantle dynamics as a pervasive in- lution of a northern Amazonian river. Geomorphology, 271, 22–­39.
fluence on Amazonian ecosystem. Scientific Reports, 9(1), 1–­11. https://doi.org/10.1016/j.geomo​rph.2016.07.030
https://doi.org/10.1038/s4159​8-­019-­53465​-­y Cronemberger, Á. A., Aleixo, A., Mikkelsen, E. K., & Weir, J. T. (2020).
Bird, J. P., Martin, R., Akçakaya, H. R., Gilroy, J., Burfield, I. J., Garnett, Postzygotic isolation drives genomic speciation between highly
S. T., Symes, A., Taylor, J., Şekercioğlu, Ç. H., & Butchart, S. H. M. cryptic Hypocnemis antbirds from Amazonia. Evolution, 74(11),
(2020). Generation lengths of the world's birds and their implica- 2512–­2525. https://doi.org/10.1111/evo.14103
tions for extinction risk. Conservation Biology, 34(5), 1252–­1261. Crouch, N. M. A., Capurucho, J. M. G., Hackett, S. J., & Bates, J. M. (2019).
https://doi.org/10.1111/cobi.13486 Evaluating the contribution of dispersal to community structure in
Borges, S. H., & Silva, J. M. C. (2012). A new area of endemism for neotropical passerine birds. Ecography, 42(2), 390–­399. https://doi.
Amazonian birds in the Rio Negro Basin. The Wilson Journal of org/10.1111/ecog.03927
Ornithology, 124(1), 15–­23. https://doi.org/10.1676/07-­103.1 Dantas, S. M., Weckstein, J. D., Bates, J., Oliveira, J. N., Catanach, T. A.,
Bouckaert, R., Vaughan, T. G., Barido-­S ottani, J., Duchêne, S., & Aleixo, A. (2021). Multi-­character taxonomic review, systemat-
Fourment, M., Gavryushkina, A., Heled, J., Jones, G., Kühnert, D., ics, and biogeography of the black-­c apped/tawny-­bellied screech
De Maio, N., Matschiner, M., Mendes, F. K., Müller, N. F., Ogilvie, owl (Megascops atricapilla-­M. watsonii) complex (Aves: Strigidae).
H. A., Du Plessis, L., Popinga, A., Rambaut, A., Rasmussen, D., Zootaxa, 4949(3), 401–­4 44.
Siveroni, I., … Drummond, A. J. (2019). BEAST 2.5: An advanced da Silva, J. M. C., & Garda, A. (2010). Padrões e Processos Biogeográficos
software platform for Bayesian evolutionary analysis. PLoS na Amazônia. In C. J. B. C. Carvalho & E. A. B. Almeida (Eds.),
Computational Biology, 15(4), 1–­28. https://doi.org/10.1371/journ​ Biogeografia da América do Sul: Padrões e Processos. Roca.
al.​p cbi.1006650 da Silva, J. M. C., Novaes, F. C., & Oren., D. C. (2002). Differentiation of
Bruen, T. C., Philippe, H., & Bryant, D. (2006). A simple and robust sta- Xiphocolaptes (Dendrocolaptidae) across the river Xingu, Brazilian
tistical test for detecting the presence of recombination. Genetics, Amazonia: Recognition of a new phylogenetic species and biogeo-
172(4), 2665–­2681. https://doi.org/10.1534/genet​ics.​105.048975 graphic implications. Bulletin of the British Ornithologists' Club, 122,
Burney, C. W., & Brumfield, R. T. (2009). Ecology predicts levels of ge- 185–­194.
netic differentiation in neotropical birds. The American Naturalist, da Silva, J. M. C., & Oren, D. C. (1996). Application of parsimony anal-
174(3), 358–­368. https://doi.org/10.1086/603613 ysis of endemicity in Amazonian biogeography: An example with
|
12      CARNEIRO et al.

primates. Biological Journal of the Linnean Society, 59(4), 427–­437. Hoorn, C., Bogotá-­A , G. R., Romero-­Baez, M., Lammertsma, E. I., Flantua,
https://doi.org/10.1006/bijl.1996.0073 S. G. A., Dantas, E. L., Dino, R., do Carmo, D. A., &Chemale JrF.
da Silva, J. M. C., Rylands, A. B., Da Fonseca, G., & a B. (2005). The fate (2016). The Amazon at sea: Onset and stages of the Amazon River
of the Amazonian areas of endemism. Conservation Biology, 19(3), from a marine record, with special reference to Neogene plant turn-
689–­694. https://doi.org/10.1111/j.1523-­1739.2005.00705.x over in the drainage basin. Global and Planetary Change, 153, 51–­65.
D'Horta, F. M., Cuervo, A. M., Ribas, C. C., Brumfield, R. T., & Miyaki, C. https://doi.org/10.1016/j.glopl​acha.2017.02.005
Y. (2013). Phylogeny and comparative phylogeography of Sclerurus Hung, C. M., Drovetski, S. V., & Zink, R. M. (2017). The roles of ecol-
(Aves: Furnariidae) reveal constant and cryptic diversification in ogy, behaviour and effective population size in the evolution of
an old radiation of rain forest understorey specialists. Journal of a community. Molecular Ecology, 26(14), 3775–­3784. https://doi.
Biogeography, 40(1), 37–­49. https://doi.org/10.1111/j.1365-­2699.​ org/10.1111/mec.14152
2012.02760.x Huson, D. H., & Bryant, D. (2006). Application of phylogenetic networks
Drummond, A. J., Ho, S. Y. W., Phillips, M. J., & Rambaut, A. (2006). in evolutionary studies. Molecular Biology and Evolution, 23(2), 254–­
Relaxed phylogenetics and dating with confidence. PLoS Biology, 267. https://doi.org/10.1093/molbe​v/msj030
4(5), e88. Irion, G., & Kalliola, R. (2010). Long-­term landscape development pro-
Drummond, A. J., & Rambaut, A. (2007). BEAST: Bayesian evolutionary cesses in Amazonia. In C. Hoorn & F. Wesselingh (Eds.), Amazonia:
analysis by sampling trees. BMC Evolutionary Biology, 7, 214. Landscape and species evolution: A look into the past (pp. 185–­197).
Fernandes, A. M., Wink, M., Sardelli, C. H., & Aleixo, A. (2014). Multiple Wiley-­Blackwell.
speciation across the Andes and throughout Amazonia: The case Isler, M. L., Bravo, G. A., & Brumfield, R. T. (2014). Systematics of the
of the spot-­backed antbird species complex (Hylophylax naevius/​ obligate ant-­following clade of antbirds ( Aves: Passeriformes:
Hylophylax naevioides). Journal of Biogeography, 41(6), 1094–­1104. Thamnophilidae ). The Wilson Journal of Ornithology, 126(4), 635–­
https://doi.org/10.1111/jbi.12277 648. https://doi.org/10.1676/13-­199.1
Ferreira, M., Aleixo, A., Ribas, C. C., & Santos, M. P. D. (2017). Biogeography Junk, W. J., Piedade, M. T. F., Schöngart, J., Cohn-­Haft, M., Adeney, J. M.,
of the neotropical genus Malacoptila (Aves: Bucconidae): The influ- & Wittmann, F. (2011). A classification of major naturally-­occurring
ence of the Andean orogeny, Amazonian drainage evolution and amazonian lowland wetlands. Wetlands, 31(4), 623–­6 40. https://
palaeoclimate. Journal of Biogeography, 44(4), 748–­759. https://doi. doi.org/10.1007/s1315​7-­011-­0190-­7
org/10.1111/jbi.12888 Rocha, D. G. da, & Kaefer, I. L. (2019). What has become of the refugia
Fu, Y. X. (1997). Statistical tests of neutrality of mutations against pop- hypothesis to explain biological diversity in Amazonia? Ecology and
ulation growth, hitchhiking and background selection. Genetics, Evolution, 9(7), 4302–­4309. https://doi.org/10.1002/ece3.5051
147(2), 915–­925. Kastner, T. P., & Goni, M. A. (2003). Constancy in the vegetation of
Haffer, J. (1969). Speciation in Amazonian forest birds. Science, 165, 131–­137. the Amazon Basin during the late Pleistocene: Evidence from
Haffer, J. (1974). Avian speciation in tropical South America (Vol. 14, pp. the organic matter composition of Amazon deep sea fan sedi-
1–­390). Nuttall Ornithol. Club. ments. Geology, 31(4), 291–­294. https://doi.org/10.1130/0091-­
Haffer, J. (1987). Biogeography of neotropical birds. In T. C. Whitmore 7613(2003)031<0291:CITVO​T>2.0.CO;2
& G. T. Prance (Eds.), Biogeography and quaternary history in tropical Katoh, K., Misawa, K., Kuma, K., & Miyata, T. (2002). MAFFT: a novel
America (pp. 105–­150). Claredon Press. method for rapid multiple sequence alignment based on fast Fourier
Haffer, J. (1992). On the “river effect” in some forest birds of southern transform. Nucleic Acids Research, 30(14), 3059–­3 066. https://doi.
Amazonia. Bol. Mus. Para. Emílio Goeldi. Série Zoologia, 8(1), 217–­246. org/10.1093/nar/gkf436
Haffer, J. (1993). Time's cycle and time's arrow in the history of Amazonia. Kearse, M., Moir, R., Wilson, A., Stones-­H avas, S., Cheung, M.,
Biogeographica, 69, 15–­45. Sturrock, S., Buxton, S., Cooper, A., Markowitz, S., Duran, C.,
Haffer, J. (2008). Hypotheses to explain the origin of species in Amazonia. Thierer, T., Ashton, B., Meintjes, P., & Drummond, A. (2012).
Brazilian Journal of Biology, 68(4), 917–­947. Geneious basic: An integrated and extendable desktop software
Harvey, M. G., Bravo, G. A., Claramunt, S., Cuervo, A. M., Derryberry, platform for the organization and analysis of sequence data.
G. E., Battilana, J., Seeholzer, G. F., Shearer, M. J., O'Meara, B. Bioinformatics, 28(12), 1647–­1649. https://doi.org/10.1093/bioin​
C., Faircloth, B. C., Edwards, S. V., Pérez-­Emán, J., Moyle, R. G., forma​t ics/bts199
Sheldon, F. H., Aleixo, A., Smith, B. T., Chesser, R. T., Silveira, L. F., Kocher, T. D., Thomas, W. K., Meyer, A., Edwards, S. V., Paabo, S.,
Cracraft, J., … Derryberry, E. P. (2020). The evolution of a tropi- Villablanca, F. X., & Wilson, A. C. (1989). Dynamics of mitochon-
cal biodiversity hotspot. Science, 370, 1343–­1348. https://doi. drial DNA evolution in animals: Amplification and sequencing with
org/10.1126/scien​ce.aaz6970 conserved primers. Proceedings of the National Academy of Sciences
Hasegawa, M., Kishino, H., &Yano, T. (1985). Dating of the human-­ape split- of the United States of America, 86(16), 6196–­6200. https://doi.
ting by a molecular clock of mitochondrial DNA. Journal of Molecular org/10.1073/pnas.86.16.6196
Evolution, 22(2), 160–­174. https://doi.org/10.1007/BF021​01694 Lanfear, R., Calcott, B., Ho, S. Y. W., & Guindon, S. (2012). PartitionFinder:
Hayakawa, E. H., & Rossetti, D. F. (2015). Late quaternary dynamics in the Combined selection of partitioning schemes and substitution mod-
Madeira river basin, southern Amazonia (Brazil), as revealed by pa- els for phylogenetic analyses. Molecular Biology and Evolution, 29(6),
leomorphological analysis. Anais Da Academia Brasileira de Ciencias, 1695–­1701.
87(1), 29–­50. https://doi.org/10.1590/0001-­37652​01520​130506 Latrubesse, E. M. (2002). Evidence of quaternary palaeohydrological
Heled, J., & Drummond, A. J. (2010). Bayesian inference of species trees changes in middle Amazonia: The Aripuanã -­Roosevelt and Jiparaná
from multilocus data. Molecular Biology and Evolution, 27(3), 570–­ “fans”. Zeitschrift für Geomorphologie, 129, 61–­72.
580. https://doi.org/10.1093/molbe​v/msp274 Latrubesse, E. M., & Rancy, A. (2000). Neotectonic influence on tropi-
Hey, J. (2010). Isolation with migration models for more than two popu- cal rivers of southwestern Amazon during the late quaternary: The
lations. Molecular Biology and Evolution, 27(4), 905–­920. https://doi. moa and Ipixuna river basins, Brazil. Quaternary International, 72(1),
org/10.1093/molbe​v/msp296 67–­72. https://doi.org/10.1016/S1040​-­6182(00)00022​-­7
Hey, J., & Nielsen, R. (2004). Multilocus methods for estimating Lerner, H. R. L., Meyer, M., James, H. F., Hofreiter, M., & Fleischer, R.
­population sizes, migration rates and divergence time, with appli- C. (2011). Report multilocus resolution of phylogeny and times-
cations to the divergence of Drosophila pseudoobscura and D. persi- cale in the extant adaptive radiation of Hawaiian honeycreepers.
milis. Genetics, 167(2), 747–­760. https://doi.org/10.1534/genet​ics.​ Current Biology, 21(21), 1838–­1844. https://doi.org/10.1016/​
103.024182 j.cub.2011.09.039
CARNEIRO et al. |
      13

Librado, P., & Rozas, J. (2009). DnaSP v5: A software for comprehensive Ricklefs, R. E., & Schluter, D. (1993). In R. E. Ricklefs & D. Schluter (Eds.),
analysis of DNA polymorphism data. Bioinformatics, 25(11), 1451–­ Species diversity in ecological communities: Historical and geographi-
1452. https://doi.org/10.1093/bioin​forma​tics/btp187 cal perspectives. University of Chicago Press.
Lynch-­Alfaro, J. W., Boubli, J. P., Paim, F. P., Ribas, C. C., Silva, M. N., Ronquist, F., & Huelsenbeck, J. P. (2003). Mrbayes 3: Bayesian phyloge-
Messias, M., Rohe, F., Merces, M. P., Silva, J. S., Jr., Silva, C. R., netic inference under mixed models. Bioinformatics, 19, 1572–­1574.
Pinho, G. M., Koshkarian, G., Nguyen, M. T., Harada, M. L., Rabelo, Roosmalen, V. M., van Roosmalen, T., Mittermeier, R. A., & Rylands, A.
R. M., Queiroz, H. L., Alfaro, M. E., & Farias, I. (2014). Biogeography B. (2000). Two new species of marmoset, genus Callithrix Erxleben,
of squirrel monkeys ( genus Saimiri ): South-­Central Amazon ori- 1777 (Callitrichidae, primates) from the Tapajós/Madeira interflu-
gin and rapid pan-­Amazonian diversification of a lowland primate. vium, south Central Amazonia, Bazil. Neotropical Primates, 8(2), 18.
Molecular Phylogenetics and Evolution, 82(B), 436–­454. https://doi. Rossetti, D. F. (2014). The role of tectonics in the late quaternary evo-
org/10.1016/j.ympev.2014.09.004 lution of Brazil's Amazonian landscape. Earth-­Science Reviews, 139,
Matzke, N. J. (2013). Probabilistic historical biogeography: New models 362–­389. https://doi.org/10.1016/j.earsc​irev.2014.08.009
for founder-­event speciation, imperfect detection, and fossils allow Soares, L. M. D. S., Bates, J., Carneiro, L. S., Santos, M. P. D., & Aleixo,
improved accuracy and model-­testing. Frontiers in Biogeography, 5, A. (2019). Molecular systematics, biogeography and taxonomy of
242–­248. forest-­falcons in the Micrastur ruficollis species complex (Aves:
Morrone, J. J. (2009). Evolutionary biogeography: An integrative approach. Falconidae). Journal of Avian Biology, 50(4), 1–­14. https://doi.
Columbia University Press. org/10.1111/jav.01943
Nielsen, R., & Wakeley, J. (2001). Distinguishing migration from isolation: Sambrook, J., & Russell, D. W. (1989). Molecular cloning: a laboratory man-
A Markov chain Monte Carlo approach. Genetics, 158(2), 885–­896. ual (Vol. 3). Cold Spring Harbor Laboratory Press.
Phillimore, A. B., Orme, C. D. L., Thomas, G. H., Blackburn, T. M., Santana, A., Silva, S. M., Batista, R., Sampaio, I., & Aleixo, A. (2020).
Bennett, P. M., Gaston, K. J., & Owens, I. P. F. (2008). Sympatric Molecular systematics, species limits, and diversification of the
speciation in birds is rare: Insights from range data and sim- genus Dendrocolaptes (Aves: Furnariidae): Insights on biotic ex-
ulations. American Naturalist, 171(5), 646–­657. https://doi. changes between dry and humid forest types in the neotropics.
org/10.1086/587074 Journal of Zoological Systematics and Evolutionary Research, 00, 1–­17.
Price, T. D., Hooper, D. M., Buchanan, C. D., Johansson, U. S., Tietze, https://doi.org/10.1111/jzs.12408
D. T., Alström, P., Olsson, U., Ghosh-­Harihar, M., Ishtiaq, F., Gupta, Schweizer, M., & Liu, Y. (2018). Avian diversity and distributions and their
S. K., Martens, J., Harr, B., Singh, P., & Mohan, D. (2014). Niche evolution through space and time. In D. Tietze (Ed.), Bird species.
filling slows the diversification of Himalayan songbirds. Nature, Fascinating life sciences. Springer. https://doi.org/10.1007/978-­3-­
509(7499), 222–­225. https://doi.org/10.1038/natur​e13272 319-­91689​-­7_8
Prychitko, T. M., & Moore, W. S. (1997). The utility of DNA sequences Silva, S. M., Townsend Peterson, A., Carneiro, L., Burlamaqui, T. C. T.,
of an intron from the β-­fibrinogen gene in phylogenetic analysis of Ribas, C. C., Sousa-­Neves, T., Miranda, L. S., Fernandes, A. M.,
woodpeckers (Aves: Picidae). Molecular Phylogenetics and Evolution, D'Horta, F. M., Araújo-­Silva, L. E., Batista, R., Bandeira, C. H. M. M.,
8(2), 193–­204. https://doi.org/10.1006/mpev.1997.0420 Dantas, S. M., Ferreira, M., Martins, D. M., Oliveira, J., Rocha, T. C.,
Pupim, F. N., Sawakuchi, A. O., Almeida, R. P., Ribas, C. C., Kern, Sardelli, C. H., Thom, G., & Aleixo, A. (2019). A dynamic continen-
A. K., Hartmann, G. A., Chiessi, C. M., Tamura, L. N., Mineli, T. D., tal moisture gradient drove Amazonian bird diversification. Science
Savian, J. F., Grohmann, C. H., Bertassoli, D. J., Stern, A. G., Cruz, Advances, 5(7), 1–­11. https://doi.org/10.1126/sciadv.aat5752
F. W., & Cracraft, J. (2019). Chronology of Terra Firme forma- Sioli, H. (1984). Hydrochemistry and geology in the Brazilian Amazon re-
tion in Amazonian lowlands reveals a dynamic quaternary land- gion. Amazoniana, 1, 74–­83.
scape. Quaternary Science Reviews, 210, 154–­163. https://doi. Slade, R. W., Moritz, C., Heidemann, A., & Hale, P. T. (1993). Rapid as-
org/10.1016/j.quasc​irev.2019.03.008 sessment of single-­copy nuclear DNA variation in diverse species.
Rambaut, A., Suchard, M., Xie, D., & Drummond, A. (2014). Tracer v. 1.6. Molecular Ecology, 2, 359–­373.
Institute of Evolutionary Biology, University of Edinburgh. Available Smith, B. T., & Klicka, J. (2010). The profound influence of the late
online at: http://Beast.Bio.Ed.Ac.Uk/Tracer Pliocene Panamanian uplift on the exchange, diversification, and
Ramos-­Onsins, S. E., & Rozas, J. (2002). Statistical properties of new distribution of New World birds. Ecography, 33, 333–­3 42. https://
neutrality tests against population growth. Molecular Biology doi.org/10.1111/j.1600-­0587.2009.06335.x
and Evolution, 19(12), 2092–­2100. https://doi.org/10.1093/oxfor​d​ Smith, B. T., Mccormack, J. E., Cuervo, A. M., Hickerson, M. J., Aleixo, A.,
jour​nals.​molbev.a004034 Cadena, C. D., Pérez-­Emán, J., Burney, C. W., Xie, X., Harvey, M. G.,
Räsänen, M. E., Salo, J. S., & Kalliola, R. J. (1987). Fluvial perturbance in Faircloth, B. C., Glenn, T. C., Derryberry, E. P., Prejean, J., Fields, S.,
the western Amazon basin: Regulation by long-­term sub-­Andean & Brumfield, R. T. (2014). The drivers of tropical speciation. Nature,
tectonics. Science, 238(4832), 1398–­1401. https://doi.org/10.1126/ 515, 406–­4 09. https://doi.org/10.1038/natur​e13687
scien​ce.238.4832.1398 Sousa-­Neves, T., Aleixo, A., & Sequeira, F. (2013). Cryptic patterns of
Ree, R. H., & Sanmartín, I. (2018). Conceptual and statistical problems diversification of a widespread Amazonian woodcreeper species
with the DEC+J model of founder-­event speciation and its compar- complex (Aves: Dendrocolaptidae) inferred from multilocus phylo-
ison with DEC via model selection. Journal of Biogeography, 45(4), genetic analysis: Implications for historical biogeography and tax-
741–­749. https://doi.org/10.1111/jbi.13173 onomy. Molecular Phylogenetics and Evolution, 68, 410–­424.
Ribas, C., Aleixo, A., Nogueira, A. C. R., Miyaki, C. Y., & Cracraft, J. Stamatakis, A. (2006). RAxML-­VI-­HPC: Maximum likelihood-­based
(2012). A palaeobiogeographic model for biotic diversification phylogenetic analyses with thousands of taxa and mixed models.
within Amazonia over the past three million years. Proceedings of Bioinformatics, 22, 2688–­2690.
the Royal Society B: Biological Sciences, 279(1729), 681–­689. https:// Stephens, M., & Donnelly, P. (2003). A comparison of bayesian meth-
doi.org/10.1098/rspb.2011.1120 ods for haplotype reconstruction from population genotype data.
Ribas, C. C., Aleixo, A., Gubili, C., D'Horta, F. M., Brumfield, R. American Journal of Human Genetics, 73(5), 1162–­1169. https://doi.
T., & Cracraft, J. (2018). Biogeography and diversification org/10.1086/379378
of Rhegmatorhina (Aves: Thamnophilidae): Implications for Tajima, F. (1989). Statistical method for testing the neutral mutation hy-
the evolution of Amazonian landscapes during the quater- pothesis by DNA polymorphism. Genetics, 123, 585–­595.
nary. Journal of Biogeography, 45(4), 917–­9 28. https://doi. Thom, G., Xue, A. T., Sawakuchi, A. O., Ribas, C. C., Hickerson, M. J.,
org/10.1111/jbi.13169 Aleixo, A., & Miyaki, C. (2020). Quaternary climate changes as
|
14      CARNEIRO et al.

speciation drivers in the Amazon floodplains. Science. Advances, 6,


1–­11. https://doi.org/10.1126/sciadv.aax4718 B I O S K E TC H
Wagenmakers, E.-­J., & Farrell, S. (2004). AIC model selection using Akaike Lincoln S. Carneiro is interested in empirical biogeography of
weights. Psychonomic Bulletin & Review, 11(1), 192–­196. https://doi. Neotropical birds, with emphasis on investigation of the evolu-
org/10.3758/BF032​06482
tionary mechanisms behind diversification of species complexes
Wallace, A. R. (1852). On the monkeys of the Amazon. Proceedings of the
in widespread biomes. All authors share a common interest in
Zoological Society of London, 20, 107–­110.
Weir, J. T., Faccio, M. S., Pulido-­Santacruz, P., Barrera-­Guzman, A. O., general evolutionary biology and biogeography.
& Aleixo, A. (2015). Hybridization in headwater regions, and the
role of rivers as drivers of speciation in Amazonian birds. Evolution, Author contributions: L.S.C., A.A., D.O. and J.M.C.S. conceived
69(7), 1823–­1834. https://doi.org/10.1111/evo.12696
the research; L.S.C., A.A. and T.C.T.B obtained most of the speci-
Willis, E. O., & Oniki, Y. (1992). As aves e as formigas de correição. Boletim
Do Museu Paraense Emilio Goeldi. Serie Zoologia, 8, 123–­150. mens; T.C.T.B. generated the genetic data; L.S.C. and T.C.T.B.
Willis, E. O. (1969). On the behavior of five species of Rhegmatorhina, ant-­ analysed the data; L.S.C. and J.M.C.S led the writing, with other
following antbirds of the Amazon basin. The Wilson Bulletin, 81(4), authors contributing; and all authors read and approved the
363–­395.
manuscript.
Willis, E. O. (1982). The behavior of scale-­backed antbirds. The Wilson
Bulletin, 94(4), 447–­462.
Willis, E. O. (1984). Phlegopsis erythroptera (Gould, 1855) and relatives (Aves, S U P P O R T I N G I N FO R M AT I O N
Formicariidae) as army ant followers. Revista Brasileira de Zoologia, Additional supporting information may be found in the online
2(3), 165–­170. https://doi.org/10.1590/s0101​-­81751​98300​0300008
version of the article at the publisher’s website.
Willis, E. O., & Oniki, Y. (1978). Birds and Army ants. Annual Review of
Ecology and Systematics, 9(1), 243–­263. https://doi.org/10.1146/
annur​ev.es.09.110178.001331
Zheng, Y., & Wiens, J. J. (2015). Do missing data influence the ac- How to cite this article: Carneiro, L., Burlamaqui, T. C.,
curacy of divergence-­time estimation with BEAST? Molecular Aleixo, A., Oren, D. C. & Silva, J. M. (2022). Biogeography and
Phylogenetics and Evolution, 85, 41–­49. https://doi.org/10.1016/j. diversification of bare-­eyes, an endemic Amazonian clade.
ympev.2015.02.002
Journal of Biogeography, 00, 1–14. https://doi.org/10.1111/
Zimmer, K., & Isler, M. L. (2020). Black-­spotted bare-­eye (Phlegopsis
nigromaculata). In E. d. J. J. del Hoyo, A. Elliott, J. Sargatal, & D. jbi.14370
A. Christie (Eds.), Birds of the world. Cornell Lab of Ornithology.
https://doi.org/10.2173/bow.bsbey​e1.01
Zink, R. M., & Barrowclough, G. F. (2008). Mitochondrial DNA under
siege in avian phylogeography. Molecular Ecology, 17(9), 2107–­2121.
https://doi.org/10.1111/j.1365-­294X.2008.03737.x

You might also like