cryptic speciation in the white shouldered antshrike thamnophilus aethiops aves thamnophilidae the...

Upload: juan-sebastian-geraldo-escobar

Post on 06-Jul-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    1/16

    Cryptic speciation in the white-shouldered antshrike (Thamnophilusaethiops, Aves – Thamnophilidae): The tale of a transcontinentalradiation across rivers in lowland Amazonia and the northeasternAtlantic Forest

    Gregory Thom a, Alexandre Aleixo b,⇑

    a Curso de Pós-graduação em Zoologia, Universidade Federal do Pará/Museu Paraense Emílio Goeldi, Caixa Postal 399, CEP 66040-170 Belém, PA, Brazilb Coordenação de Zoologia, Museu Paraense Emílio Goeldi, Caixa Postal 399, CEP 66040-170 Belém, PA, Brazil

    a r t i c l e i n f o

     Article history:

    Received 23 June 2014

    Revised 20 September 2014

    Accepted 26 September 2014

    Available online 5 October 2014

    Keywords:

    Diversification hypotheses

    Historical biogeography

    Phylogeography

    Population genetics

    Refuges

    Taxonomy

    a b s t r a c t

    The growing knowledge on paleogeography and the recent applications of molecular biology and phylo-

    geography to the study of the Amazonian biota have provideda framework for testing competinghypoth-

    eses of biotic diversification in this region. Here, we reconstruct the spatio-temporal context of 

    diversification of a widespread understory polytypic Amazonian bird species (Thamnophilus aethiops)

    and contrast it with different hypotheses of diversification and the taxonomy currently practiced in the

    group. Sequences of mtDNA (cytochrome  b  and ND2) and nuclear (b-fibrinogen introns 5 and 7 and the

    Z-liked Musk4) genes, addingup to 4093 bp of 89 individuals covering the Amazonian, Andean, andAtlan-

    tic Forest populations of   T. aethiops   were analyzed. Phylogenetic and population genetics analyses

    revealed ten reciprocally monophyletic and genetically isolated or nearly-isolated lineages in  T. aethiops,

    highlighting several inconsistencies between taxonomy and evolutionary history in this group. Our data

    suggest that the diversification of  T. aethiops  started in the Andean highlands, and then proceeded into

    the Amazonian lowlands probably after the consolidation of the modern Amazonian drainage. The main

    cladogenetic events in T. aethiops may be related to the formation and structuring of large Amazonian riv-

    ers during the Late Miocene–Early Pleistocene, coinciding with the dates proposed for other lineages of 

    Amazonian organisms. Population genetics data do not support climatic fluctuations as a major source

    of diversification in   T. aethiops. Even though not entirely concordant with paleobiogeographic models

    derived from phylogenies of other vertebrate lineages, our results support a prominent role for rivers as

    major drivers of diversification in Amazonia, while underscoring that different diversification scenarios

    are probably related to the distinct evolutionary origins of groups being compared.

     2014 Elsevier Inc. All rights reserved.

    1. Introduction

    Since the 19th century, the high species diversity of the Amazon

    allied to multifaceted patterns of geographical distributions insti-gated naturalists and researchers to explain such complexity

    (Wallace, 1852; Haffer, 1969; Bates, 2001).

    Despite the large number of distinct hypotheses to explain

    Amazonian biodiversity (e.g., refuge hypothesis, Haffer, 1969; riv-

    erine hypothesis,   Wallace, 1852; Ayres and Clutton-Brock, 1992;

    riverine-refuge hypothesis,   Ayres and Clutton-Brock, 1992;

    Haffer, 1993, 2001; ecological gradients hypothesis,  Endler, 1977;

    ‘‘museum’’ hypothesis,   Roy et al., 1997; and marine incursions,

    Bates, 2001), and the common sense that many causations have

    operated for the formation of such diversity (Bush, 1994; Haffer,

    2001; Miller et al., 2008), recent phylogeographic and paleobioge-

    ographic reconstructions (Aleixo and Rossetti 2007; Patel et al.,2011; Weir and Price, 2011; Ribas et al., 2012) have postulated

    the formation of the current Amazonian physical landscape as

    the main source of cladogenetic events among the studied lineages.

    The formation of the Amazon basin was ultimately driven by

    the Andean uplift and related arches, which resulted in the forma-

    tion of a large fluvio-lacustrine system in western Amazonia, dur-

    ing the early Miocene (Espurt et al., 2010; Mora et al., 2010). With

    the continuous rise of the Andes in the middle Miocene, this sys-

    tem probably expanded to the Purus Arch, a tectonic structure

    located 300 km west of Manaus, and which separates the Solimões

    and Amazonas sedimentary basins (Figueiredo et al., 2009; Hoorn

    http://dx.doi.org/10.1016/j.ympev.2014.09.023

    1055-7903/  2014 Elsevier Inc. All rights reserved.

    ⇑ Corresponding author.

    E-mail address: [email protected] (A. Aleixo).

    Molecular Phylogenetics and Evolution 82 (2015) 95–110

    Contents lists available at   ScienceDirect

    Molecular Phylogenetics and Evolution

    j o u r n a l h o m e p a g e :   w w w . e l s e v i e r . c o m / l o c a t e / y m p e v

    http://dx.doi.org/10.1016/j.ympev.2014.09.023mailto:[email protected]://dx.doi.org/10.1016/j.ympev.2014.09.023http://www.sciencedirect.com/science/journal/10557903http://www.elsevier.com/locate/ympevhttp://www.elsevier.com/locate/ympevhttp://www.sciencedirect.com/science/journal/10557903http://dx.doi.org/10.1016/j.ympev.2014.09.023mailto:[email protected]://dx.doi.org/10.1016/j.ympev.2014.09.023http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://crossmark.crossref.org/dialog/?doi=10.1016/j.ympev.2014.09.023&domain=pdfhttp://-/?-

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    2/16

    et al., 2010). The period when the Purus Arch was breached prob-

    ably marked the onset of the development of the transcontinental

    Amazon River, draining eastward the western fluvio-lacustrine

    system, and enabling the formation of upland  terra-firme   forests

    in the western Amazonia; however, the timing of this event is still

    a matter of controversy, with two main proposed periods based on

    different sources of evidence: late Miocene, between 6.5 and 10 Ma

    (Hoorn et al., 2010; Latrubesse et al., 2010) and Pliocene (approx-imately 2.5 Ma; Campbell et al., 2006). Despite the fact that major

    interfluves of the Amazonian region delineate the geographical

    distribution of most endemic vertebrate taxa and that this has pro-

    vided the basis for the recognition of the so called areas of ende-

    mism (Cracraft, 1985; Silva et al., 2005), the process of river

    dynamics driving population divergence is still poorly understood,

    with multiple area-relationships patterns documented so far

    (Aleixo, 2004; Fernandes et al., 2012; Ribas et al., 2012; d’Horta

    et al., 2013). This suggests that although rivers have been barriers

    to gene flow, other historical events probably contributed to the

    overall diversification pattern. However, this knowledge is still

    meager when compared with the rich and complex biodiversity

    of Amazonia, preventing the recognition of a general model or sets

    of heuristic models of diversification, corroborated by several

    lineages.

    The polytypic species   Thamnophilus aethiops   (Aves: Thamno-philidae) is a good model to study the paleobiogeography of the

    Amazon basin due the following reasons: (1) it is widely distrib-

    uted in Amazonia and neighboring areas such as the Andes and

    the Atlantic Forest in eastern Brazil, two areas known to be histor-

    ically connected to the development of modern Amazonia; (2) it is

    restricted to the understory of upland  terra-firme   forest, rapidlyresponding to environmental change (Stotz et al., 1996); and (3)

    subspecies distributions are mainly bounded by the major tributar-

    ies of the Amazon River, with some important exceptions that can

    provide insights into the circumstances whereby a single lineage

    does and does not respond to rivers as barriers. Thus, we estimated

    the spatio-temporal scenario of diversification of the polytypic  T.

    aethiops to address the following questions: (1) what are the evo-lutionary relationships among the populations/subspecies of   T.aethiops? (2) what are the relationships between the events that

    led to diversification in T. aethiops and the paleogeographic modelsproposed for the formation of Amazonia? (3) do the demographic

    history of the studied populations support any particular diversifi-

    cation hypothesis?

    2. Material and methods

     2.1. Samples, laboratory procedures, and data analyses

    A total of 89 individuals from 55 localities were sampled

    throughout the distribution of  T. aethiops in Amazonia, northeast-ern Atlantic Forest and the Andean foothills, covering nine of ten

    described subspecies (Zimmer and Isler, 2003;   Fig. 1b,  Table 1).

    Only  T. a. wetmorei, purportedly endemic to the Andean foothills

    of Colombia, was not sampled. Nonetheless, we assume here that

    that this taxon is closely related to the  polionotus subspecies fromnorthwestern Amazonia, from which it is hardly differentiated

    based on plumage characters (Zimmer and Isler, 2003). Subspecific

    identification of our samples were made through the inspection of 

    voucher specimens and followed the most recent taxonomy pro-

    posed for T. aethiops (Zimmer and Isler 2003). We used Thamnophi-lus aroyae,   T. unicolor , and   T. caerulescens  as outgroups due theirclose relationship to T. aethiops  (Brumfield and Edwards, 2007).

    Total DNA was extracted from approximately 20 mg of muscle

    tissue following a standard phenol/chloroform protocol(Sambrook et al., 1989) or with the aid of a DNeasy Quiagen

    (Hilden, Germany) extraction kit. We amplified the mitochondrial

    genes cytochrome  b  (cyt  b, 992 bp,  n  = 88) and NADH dehydroge-nase subunit 2 (ND2, 1041 bp, n = 79), as well as the nuclear genesb-fibrinogen intron 7 (Bf7, 963 bp,  n = 51),   b-fibrinogen intron 5

    (Bf5, 532 bp, n = 63), and the intron 4 of the skeletal muscle recep-tor of tyrosine kinase (Musk4, 565 bp, n = 64) linked to the Z chro-mosome (see   Table 2   for the primers used). Polymerase chain

    reaction (PCR) was performed with 25ll of final volume, andapproximately 50 ng of genomic DNA, 1.5–2.5 mM of MgCl2,200 mM of dNTPs and 0.1 U of Taq DNA polymerase Promega

    (Madison, WI, USA). Reactions started with a denaturation step at

    94 C for 5 min, followed by 35 cycles of three steps: (1) 94 C for

    1 min; (2) annealing temperatures ranging from 50 C (cyt  b  andBf5) and 52 C (Musk4) to 59 C (ND2, Bf7) for 1 min; and (3)

    72 C for 1 min. The last step, for the extension, was at 72 C for

    5 min. PCR products were purified with the Polyethylene glycol

    protocol (PEG). Sequencing was carried out on an ABI 3130 auto-

    mated capillary sequencer (Applied Biosystems, Foster City, Cali-

    fornia, USA) with the ABI Prism Big Dye terminator Kit. To

    confirm observed mutations, both strands of each sample were

    sequenced. The DNA strands were edited and aligned manually

    in BIOEDIT 7.0.5 (Hall, 1999). Saturation of the nucleotide substitu-

    tions in the mitochondrial DNA were evaluated using the software

    DAMBE (Xia and Xie, 2001). For nuclear loci, heterozygous nucleo-

    tide positions were inferred by the presence of double peaks of the

    same size in the electropherogram. To obtain the gametic phase of 

    haplotypes of the heterozygous individuals, we used a Bayesian

    approach as implemented in PHASE 2.1 1 (Stephens et al., 2001;

    Stephens and Donnelly, 2003; Stephens and Scheet, 2005). The

    threshold of 70% of posterior probability was assumed as a confi-

    dence value for the analyzed haplotypes (see   Harrigan et al.,

    2008). Lower values were regarded as ambiguous. To obtain the

    gametic phase for individuals heterozygous in size (presence of 

    indels in one of the strands), we used the program CHAMPURU

    v1.0 (Flot et al., 2006; Flot, 2007; available online at   http://

    www.mnhn.fr/jfflot/champuru/). Females did not have the gametic

    phase estimated for Musk4, since this gene is linked to de Z chro-mosome. To test the hypothesis of neutral evolution among inde-

    pendent loci we performed the Tajima’s D test (Tajima, 1989)

    using DNASP v.4.10.9 (Rozas et al., 2003). The significance for these

    tests was assessed through 10.000 coalescence simulations. To

    check for possible recombination among the nuclear loci we used

    the   phi   test as implemented in the SPLITSTREE 4.10 software

    (Huson and Bryant, 2006).

     2.2. Phylogenetic analyses

     2.2.1. Concatenated genes and gene treesThe phylogenetic analyses were performed using Bayesian

    inference (BI) in MrBAYES 3.1.2 (Ronquist and Huelsenbeck,

    2003) and Maximum likelihood (ML) in RaxML 7.0.3 (Stamatakis,2006). The evolutionary models which best explain the evolution

    of each dataset were selected in JMODELTEST 0.1.1 (Posada,

    2008), using the Bayesian information criterion (BIC) for BI and

    the Akaike information criterion (AIC) for ML. For the BI analyses

    with the concatenated dataset, a Bayes factor analysis (Brandley

    et al., 2005) selected four partitions (mtDNA + one separate parti-

    tion for each nuclear gene) as the best partitioning scheme. The

    evolutionary models for each partition were: (1) mtDNA (GTR + G);

    (2) Bf5 (GTR + I + G); (3) Bf7 (GTR + I); and (4) Musk4 (HKY + G).

    The BI estimated based on mtDNA only used one model for the

    two genes (GTR + G). BI analyses were generated through two

    independent runs of 1 107 generations, each with four Markov

    chains. The parameters of the chains were sampled every 1000

    generations and the first 1000 trees were discarded as burn-in.The posterior probabilities for each estimated node were obtained

    96   G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110

    http://www.mnhn.fr/jfflot/champuru/http://www.mnhn.fr/jfflot/champuru/http://www.mnhn.fr/jfflot/champuru/http://www.mnhn.fr/jfflot/champuru/

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    3/16

    Fig. 1.  (a) Phylogenetic relationships among major clades of the polytypic  Thamnophilus aethiops and outgroups derived from Bayesian and Maximum likelihood inferences

    based on sequences of the mtDNA genes (cyt b and ND2; 2033 bp) and all loci combined (cyt b, ND2, Bf5, Bf7 and Musk4; 4093 bp). Nodal support values correspond to

    posterior probabilities (above) and bootstrap pseudo-replicates (below) obtained, respectively, in the mtDNA and concatenated dataset analyses. (b) Map showing the current

    distribution and sampled localities of major clades of the polytypic  T. aethiops. Sample codes are the same used in Fig. 1a and Table 1. (c) Haplotype network of the five

    sequenced genes estimated through HaploViewer (Salzburger et al., 2011). Clade colors are the same as in Fig. 1a and 1b. Numbers inside circles represent actual number of 

    individuals having a particular haplotype, with circle sizes being proportional to it, while numbers on connecting branches are nucleotide substitutions higher than one

    separating haplotypes. Asterisks beside light blue circles represent the number of individuals from the Pernambuco area of endemism. (d) Bayesian Phylogenetics and

    Phylogeography (BP&P) results. Numbers on braches represent the number of times a given node was recovered with a speciation probability P0.95 based on six analyses

    performed using two algorithms (0 and 1) and three combinations of priors on ancestral population sizes and divergence times (see methods). (e) Overlapped Bayesian

    estimatedchronogramsderived fromalternativedating analyses based on theconcatenated mtDNA (cyt b andND2; 2033 bp;branches and95% of confidence interval in dark

    gray) and a multi-locus coalescent species tree (mtDNA, Bf5/Bf7 andMusk4; 4093 bp; branches and 95%confidence intervals in Black). Asterisks andcrosses represent values

    of posterior probability P95% for the species tree and concatenated mtDNA analyses, respectively. (For interpretation of the references to color in this figure legend, the

    reader is referred to the web version of this article.)

    G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110   97

    http://-/?-

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    4/16

     Table 1

    Code (asin Fig. 1a andb), taxon(speciesor subspecies followingthe taxonomy in Zimmer and Isler, 2003), lineage (namedaccordingto recognizedNeotropicalareas of endemism as

    in Silvaet al., 2005, withsome modifications), N – vouchernumber, Inst.– institution(MPEG: MuseuParaense Emílio Goeldi; LSUMZ: Louisiana State UniversityMuseum of Natural

    Science), and locality information (presented coordinates are plotted in  Fig. 1b) associated with each sample of  Thamnophilus aethiops and outgroups sequenced in this study.

    Code Taxon Lineage N   Inst. Locality Coordinates

    T1   T. a. atriceps   Tapajós 63,888 MPEG Brazil, Pará, Altamira, Floresta Nacional de Altamira 06S0401200

    55W1901200

    T2   T. a. atriceps   Tapajós 70,861 MPEG Brazil, Pará, Itaituba, Miritituba 04S2602400

    55W380

    2400

    T2   T. a. atriceps   Tapajós 70,862 MPEG Brazil, Pará, Itaituba, Miritituba

    T3   T. a. atriceps   Tapajós 59,159 MPEG Brazil, Pará, Novo Progresso, 20 km SW 07S1102400

    55W3000000

    T3   T. a. atriceps   Tapajós 59,160 MPEG Brazil, Pará, Novo Progresso, 20 km SW

    T3   T. a. atriceps   Tapajós 59,161 MPEG Brazil, Pará, Novo Progresso, 20 km SW

    X1   T. a. atriceps   Xingu 57,707 MPEG Brazil, Pará, Senador José Porfírio, Rio Xingu 02S3202400

    51W3300000

    X1   T. a. atriceps   Xingu 57,708 MPEG Brazil, Pará, Senador José Porfírio, Rio Xingu

    X1   T. a. atriceps   Xingu 57,709 MPEG Brazil, Pará, Senador José Porfírio, Rio Xingu

    X2   T. a. atriceps   Xingu 70,635 MPEG Brazil, Pará, Carajás, Serra dos Carajás 06S0000000

    51W1904800

    X3   T. a. atriceps   Xingu 70,658 MPEG Brazil, Pará, Carajás, Serra dos Carajás, Serra Norte 06S0900000

    50W2400000

    P1   T. a. distans   Pernambuco 70,454 MPEG Brazil, Alagoas, Ibateguara, Engenho Coimbra 09S1001200

    35W3202400

    P1   T. a. distans   Pernambuco 70,455 MPEG Brazil, Alagoas, Ibateguara, Engenho Coimbra

    P1   T. a. distans   Pernambuco 70,457 MPEG Brazil, Alagoas, Ibateguara, Engenho Coimbra

    P2   T. a. distans   Pernambuco 70,459 MPEG Brazil, Pernambuco, Barreiros, Cachoeira Linda 08S4800000

    35W1901200

    P2   T. a. distans   Pernambuco 70,460 MPEG Brazil, Pernambuco, Barreiros, Cachoeira Linda

    P2   T. a. distans   Pernambuco 70,458 MPEG Brazil, Pernambuco, Barreiros, Cachoeira Linda

    B1   T. a. incertus   Belém 58,647 MPEG Brazil, Pará, Santa Bárbara, Parque Ecológico GUNMA 01S3303600

    47W3903600

    B2   T. a. incertus   Belém 61,190 MPEG Brazil, Pará, Barcarena 02S1800000

    48W3903600

    B2   T. a. incertus   Belém 61,192 MPEG Brazil, Pará, Barcarena

    B3   T. a. incertus   Belém 58,632 MPEG Brazil, Pará, Tomé-Açú, Quatro Bocas 02S3202400

    48W0104800

    B4   T. a. incertus   Belém 70,559 MPEG Brazil, Pará, Curuça 01S0202400

    47W2703600

    EI1   T. a. injunctus   Eastern Inambari 68,883 MPEG Brazil, Amazonas, Careiro, Br 319 km 04S0404800

    60W3903600

    EI2   T. a. injunctus   Eastern Inambari 71,054 MPEG Brazil, Amazonas, Humaitá, Rio Ipixuna 07S3101200

    63W2002400

    EI2   T. a. injunctus   Eastern Inambari 71,055 MPEG Brazil, Amazonas, Humaitá, Rio Ipixuna

    EI2   T. a. injunctus   Eastern Inambari 71,056 MPEG Brazil, Amazonas, Humaitá, Rio Ipixuna

    EI2   T. a. injunctus   Eastern Inambari 71,057 MPEG Brazil, Amazonas, Humaitá, Rio Ipixuna

    EI3   T. a. injunctus   Eastern Inambari 60,637 MPEG Brazil, Acre, Senador Guiomard, Ramal Nabor Júnior 09S4604800

    67W1904800

    WI1   T. a. kapouni   Western Inambari 58,016 MPEG Brazil, Acre, Parque Nacional Serra do Divisor 08S2100000

    73W1803600

    WI2   T. a. kapouni   Western Inambari 59,837 MPEG Brazil, Acre, Assis Brasil, ESEC Rio Acre 11S0303600

    70W1601200

    WI3   T. a. kapouni   Western Inambari 60,171 MPEG Brazil, Amazonas, RDS Cujubim, Rio Jutaí 04S3900000

    68W1904800

    WI4   T. a. kapouni   Western Inambari 60,631 MPEG Brazil, Acre, Tarauacá, Br 364 km 40, Rio Liberdade 07S5302400

    71W3903600

    WI4   T. a. kapouni   Western Inambari 60,632 MPEG Brazil, Acre, Tarauacá, Br 364 km 40, Rio Liberdade

    WI5   T. a. kapouni   Western Inambari 62,071 MPEG Brazil, Acre, Porto Walter, Igarapé Cruzeiro do Vale 08S3300000

    72W5201200

    WI6   T. a. kapouni   Western Inambari 40,722 LSUMZ Peru, Loreto, 86 km SE Juanjui 07S3301900

    75W5403200

    WI6   T. a. kapouni   Western Inambari 40,661 LSUMZ Peru, Loreto, 86 km SE Juanjui

    WI7   T. a. kapouni   Western Inambari 11,235 LSUMZ Peru, Ucayali, SE Cerro Tahuayo, ENE Pucallpa 08S2503300

    74W2202600

    WI8   T. a. juruanus   Western Inambari 59,966 MPEG Brazil, Acre, Rio Branco, Fazenda Experimental Catuaba 02S1904800

    67W3600000

    WI9   T. a. juruanus   Western Inambari 60,634 MPEG Brazil Acre, Bujari, Floresta Estadual do Antimary 09S2100000

    68W0600000

    WI9   T. a. juruanus   Western Inambari 60,636 MPEG Brazil Acre, Bujari, Floresta Estadual do Antimary

    WI10   T. a. juruanus   Western Inambari 60,638 MPEG Brazil, Acre, Plácido de Castro, Ramal Novo Horizonte 10S0704800

    67W1904800

    WI11   T. a. juruanus   Western Inambari 61,297 MPEG Brazil, Acre, Rio Branco, Ramal Jarinal 09S5403600

    68W1901200

    WI11   T. a. juruanus   Western Inambari 61,298 MPEG Brazil, Acre, Rio Branco, Ramal Jarinal

    WI11   T. a. juruanus   Western Inambari 61,300 MPEG Brazil, Acre, Rio Branco, Ramal Jarinal

    98   G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    5/16

     Table 1 (continued)

    Code Taxon Lineage N   Inst. Locality Coordinates

    WI12   T. a. juruanus   Western Inambari 62,069 MPEG Brazil, Acre, Mâncio Lima, Estrada do Barão 07S3300000

    72W3603600

    WI12   T. a. juruanus   Western Inambari 62,070 MPEG Brazil, Acre, Mâncio Lima, Estrada do Barão

    WI13   T. a. juruanus   Western Inambari 59,965 MPEG Brazil, Acre, Porto Acre, Reserva Humaitá 09S4503600

    67W4001200

    WI14   T. a. juruanus   Western Inambari 63,336 MPEG Brazil, Acre, Santa Rosa, Rio Purus 09S0701200

    69W4904800

    WI15   T. a. juruanus   Western Inambari 64,546 MPEG Brazil, Acre, Senador Guiomard, Ramal Oco do Mundo 09S5002400

    67W4901200

    WI16   T. a. juruanus   Western Inambari 8984 LSUMZ Bolivia, Pando, Nicolás Suarez, 12km by road of Cobija 11S1000200

    68W2601800

    WI17   T. a. juruanus   Western Inambari 1164 LSUMZ Bolivia, La Paz, 20 k m by Rio Beni N. Puerto Linares 15S1400600

    67W3701100

    EN1   T. a. polionotus   Eastern Negro 59,521 MPEG Brazil, Amazonas, Barcelos, Rio Aracá 00S2501200

    62W5602400

    EN1   T. a. polionotus   Eastern Negro 59, 522 MPEG Brazil, Amazonas, Barcelos, Rio Aracá

    EN1   T. a. polionotus   Eastern Negro 59, 523 MPEG Brazil, Amazonas, Barcelos, Rio Aracá

    WN1   T. a. polionotus   Western Negro 59,524 MPEG Brazil, Amazonas, Barcelos, Rio Cuiuni 00S4604800

    63W1604800

    WN2   T. a. polionotus   Western Negro 59,525 MPEG Brazil, Amazonas, Novo Airão, Igarapé-Açu 02S5100000

    60W5100000

    WN3   T. a. polionotus   Western Negro 62,770 MPEG Brazil, Amazonas, Maraã, Lago Cumapi 01S4304800

    65W520

    4800

    WN3   T. a. polionotus   Western Negro 62, 771 MPEG Brazil, Amazonas, Maraã, Lago Cumapi

    NR1   T. a. punctuliger    Northern Rondônia 62,215 MPEG Brazil, Amazonas, Juruti, Base Capiranga 02S2801200

    56W0000000

    NR1   T. a. punctuliger    Northern Rondônia 62,216 MPEG Brazil, Amazonas, Juruti, Base Capiranga

    NR1   T. a. punctuliger    Northern Rondônia 58,100 MPEG Brazil, Amazonas, Juruti, Base Capiranga

    NR2   T. a. punctuliger    Northern Rondônia 68,944 MPEG Brazil, Pará, Jacareacanga, Terra do Burandir 06S1500000

    57W5201200

    NR2   T. a. punctuliger    Northern Rondônia 68,965 MPEG Brazil, Pará, Jacareacanga, Terra do Burandir

    NR3   T. a. punctuliger    Northern Rondônia 66,127 MPEG Brazil, Pará, Santarém, Retiro 02S2400000

    55W4702400

    NR4   T. a. punctuliger    Northern Rondônia 67,048 MPEG Brazil, Amazonas, Maués, Flona do Pau Rosa 03S5403600

    58W2400000

    SR1   T. a. punctuliger    Southern Rondônia 54,953 MPEG Brazil, Rondônia, Guajará-Mirim, REBIO Ouro Preto 10S4904800

    64W4500000

    SR1   T. a. punctuliger    Southern Rondônia 54,955 MPEG Brazil, Rondônia, Guajará-Mirim, REBIO Ouro Preto

    SR2   T. a. punctuliger    Southern Rondônia 61, 575 MPEG Brazil, Mato Grosso, Querência, Fazenda Tanguro 12S5302400

    52W2201200

    SR2   T. a. punctuliger    Southern Rondônia 61,576 MPEG Brazil, Mato Grosso, Querência, Fazenda Tanguro

    SR2   T. a. punctuliger    Southern Rondônia 70,226 MPEG Brazil, Mato Grosso, Querência, Fazenda Tanguro

    SR3   T. a. punctuliger    Southern Rondônia 67,397 MPEG Brazil, Mato Grosso, Paranaíta, Rio Teles Pires 09S3300000

    56W3401200

    SR3   T. a. punctuliger    Southern Rondônia 67,396 MPEG Brazil, Mato Grosso, Paranaíta, Rio Teles Pires

    SR3   T. a. punctuliger    Southern Rondônia 67,398 MPEG Brazil, Mato Grosso, Paranaíta, Rio Teles Pires

    SR4   T. a. punctuliger    Southern Rondônia 67, 442 MPEG Brazil, Mato Grosso, Paranaíta, Fazenda Aliança 09S4004800

    56W5103600

    SR4   T. a. punctuliger    Southern Rondônia 67,443 MPEG Brazil, Mato Grosso, Paranaíta, Fazenda Aliança

    SR5   T. a. punctuliger    Southern Rondônia 69, 215 MPEG Brazil, Mato Grosso, Paranaíta, Fazenda Paranaíta 09S3404800

    56W4301200

    SR6   T. a. punctuliger    Southern Rondônia 58,704 MPEG Brazil, Amazonas, Humaitá, T. Indígena Parintintin 07S5700000

    62W2201200

    SR6   T. a. punctuliger    Southern Rondônia 58,705 MPEG Brazil, Amazonas, Humaitá, T. Indígena Parintintin

    SR6   T. a. punctuliger    Southern Rondônia 58,707 MPEG Brazil, Amazonas, Humaitá, T. Indígena Parintintin

    SR7   T. a. punctuliger    Southern Rondônia 14,649 LSUMZ Bolivia, Santa Cruz, Serrania de Huanchaca 14S3104800

    60W410

    2400

    A1   T. a. aethiops   5500 LSUMZ Peru, San Martín, 28 k m by road NE Tarapoto 06S2600600

    76W1900400

    T aroyae   38,993 LSUMZ Bolivia, Cochabamba, Chapare

    T aroyae   38,985 LSUMZ Bolivia, Cochabamba, Chapare

    T aroyae   22,836 LSUMZ Bolivia, La Paz

    T. unicolor    12,144 LSUMZ Ecuador, Pichincha, Mindo

    T. unicolor    6196 LSUMZ Morona-Santiago Province

    T. unicolor    7741 LSUMZ Ecuador, Caóar

    T. unicolor    32,602 L SUMZ P eru, Cajamarca

    T. unicolor    43,515 LSUMZ Peru, San Martín

    G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110   99

    http://-/?-

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    6/16

    through a majority rule consensus of the remaining MCMC sam-

    ples. The ML analyses were not partitioned and the evolutionary

    models selected were GTR + I + G for the concatenated (all loci)

    and GTR + G for the mtDNA datasets. Nodal support was estimated

    with 1000 bootstrap pseudo-replicates. Both BI and ML analyses

    with the concatenated dataset (all loci) were performed using

    the nuclear genes with both gametic phases and the mitochondrial

    genes duplicated. To visualize genealogical relationships among

    individuals of each sequenced gene, haplotype networks were con-

    structed in HaploViewer (Salzburger et al., 2011) based on the

    topologies recovered in the species tree analysis. Mean pairwise

     p-distances (Nei, 1987) within and among populations were calcu-lated using the mtDNA dataset only in MEGA 4.0 ( Tamura et al.,

    2007).

     2.2.2. Species treeRecent approaches have shown that multi-locus analyses with

    concatenated datasets can result in well-supported topologies thatare incongruent with the true species tree, particularly in recent

    scenarios of diversification (Degnan and Rosenberg, 2009;

    Kubatko and Degnan, 2007). Thus, we used a Bayesian hierarchical

    model implemented in   *BEAST 1.6.1 (Drummond and Rambaut,

    2007; Heled and Drummond, 2010) to obtain an estimate of the

    species tree for the different lineages of   T. aethiops. For thisapproach, we assumed three independent loci: (1) mtDNA (cyt  band ND2); (2) Bf5 and Bf7; and (3) Musk4. The evolutionary models

    were the same used in the BI for the concatenated dataset, with the

    addition of the locus joining the Bf5 and Bf7 genes, for which the

    GTR + I + G model was selected. We performed an initial run of 

    5 107 generations to optimize the analysis operators. Afterwards,

    five independent runs of 5 107 generations were performed,

    sampling the parameters of the Markov chain every 1000 genera-

    tions, until the effective sample sizes (ESS) of all parameters were

    equal or higher than 200 (Kuhner 2008). Four independent runs

    were performed and combined in a majority rule consensus tree

    with posterior probability support. TRACER 1.5 (Rambaut and

    Drummond, 2007) was used to access ESS values and verify when

    the MCMC reached stable and convergent values, allowing for a

    definition of burn-in values. Based on the obtained results and

    under a conservative approach, the first 5000 samples of each

    run were discarded as burn-in. The purported species in this anal-

    ysis were the reciprocally monophyletic populations of  T. aethiopsrevealed by the BI and ML analyses (Fig. 1a).

     2.2.3. Species delimitation

    We implemented a coalescent-based bayesian modelingapproach to generate speciation probabilities of closely related

    taxa from multi locus sequence data using the program Bayesian

    Phylogenetics and Phylogeography (BP&P v.2.0b;   Yang and

    Rannala 2010), considered the most accurate coalescent-based

    method for species delimitation evaluated by   Camargo et al.

    (2012). The model implemented in BP&P assumes no gene flow

    among species, no recombination, and takes into account gene tree

    uncertainty and lineage sorting. BP&P incorporates a model that

    include the species divergence times (s), and the population sizeparameters h = 4N l, where  N  is the effective population size andl is the substitution rate per site per generation. However, reason-able prior distributions of population parameters can be difficult to

    determine, potentially affecting the posterior probability of species

    delimitations (Yang and Rannala 2010). Thus, we implemented the

    approach of  Leaché and Fujita (2010), performing analyses using

    three combination of priors representing different population sizes

    and ages for the root of the species tree as follows: (1) large ances-

    tral population size and deep divergences (h  and  s  gamma priorsG(1, 10) and G(1, 10)); (2) small ancestral population size and shal-

    low divergences among species (h and s  gamma priors G(2, 2000)and G(2, 2000)); and 3) large ancestral population size and shallow

    divergences among species (h   and  s   gamma priors G(1, 10) andG(2, 2000)). We ran the analyses for 1 106 generations, sampling

    every five and discarding the first 5 104 generations as burn-in,

    using the algorithms 0 and 1 with different fine tuning parameters

    (e = 5 for algorithm 0 and a = 2 and m = 1 for algorithm 1). We usedthe populations and the topology obtained through the species tree

    analyses as the guide tree (Fig. 1e). The population occurring dis-

     junctly in the Pernambuco region (Atlantic Forest) was considered

    a distinct lineage, sister to that of the Belém area of endemism

    population.

     2.3. Molecular dating 

    To estimate the divergence times among T. aethiops populationsrevealed by the BI and ML phylogeny estimates, a second BI was

    carried out using only the mtDNA dataset (cyt  b   and ND2). Afterthe selection of the model of molecular evolution for this dataset

    with JMODELTEST 0.1.1 (Posada, 2008), a preliminary run of 

    1 107 generations was performed in BEAST 1.6.1 (Drummond

    and Rambaut, 2007) fixing the rate of nucleotide mutations at

    1.0 per millions years for the cyt  b  (for which a robust calibrationis available; Weir and Schluter, 2008) and estimating that for the

    ND2 under the options relaxed clock and uncorrelated lognormal.

    This analysis was performed to evaluate whether both mitochon-

    drial genes were evolving at similar rates and the cyt b  calibration

    could be applied to ND2. After the confirmation that these twogenes were evolving at similar rates, the overall molecular clock

     Table 2

    List of primers used in PCR amplifications and in DNA cycles sequencing reactions of samples of  Thamnophilus aethiops   and outgroups sequenced in this study.

    Gene Primer Sequence 50–30 Source

    Cyt b   L14990 CCA TCC AAC ATC TCA GCA TGA TGA AA   Brumfield et al. (2007)

    H16065 AAC TGC AGT CAT CTC CGG TTT ACA AGA C   Brumfield et al. (2007)

    ND2 L5215 TAT CGG GCC CAT ACC CCG AAA AT   Brumfield et al. (2007)

    H5677 DGA DGA RAA DGC YAR RAY YTT DCG   Brumfield et al. (2007)

    L5758 GGN GGN TGR RBH GGN YTD AAY CAR AC   Brumfield et al. (2007)

    H6313 CTC TTA TTT AAG GCT TTG AAG GC   Brumfield et al. (2007)

    bf7 Fib-B17L TCC CCA GTA GTA TCT GCC ATT AGG GTT   Prychitko and Moore (1997)

    Fib-B17U GGA GAA AAC AGG ACA ATG ACA ATT CAC   Prychitko and Moore (1997)

    bf5 Fib5L CGC CAT ACA GAG TAT ACT GTG ACA T   Brumfield et al. (2007)

    Fib5H GCC ATC CTG GCG ATC TGA A   Brumfield et al. (2007)

    Musk4 Musk13R CTCTGAACATTGTGGATCCTCAA   Clark and Witt (2)006

    Musk13F CTTCCATGCACTACAATGGGAAA   Clark and Witt (2006)

    100   G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    7/16

    calibration used was 2.1% of nucleotide substitutions per million

    years (Weir and Schluter, 2008), under the relaxed clock and

    uncorrelated lognormal options. We performed an initial run of 

    1 107 generations to optimize the analysis operators. Afterwards,

    two independent runs of 1 107 generations were performed,

    sampling the parameters of the Markov chain every 1000 genera-

    tions, until the ESS of all parameters were equal or higher than

    200. Based on the obtained results and under a conservativeapproach, the first 800 samples of each run were discarded as

    burn-in. Values of posterior probability and divergence times were

    estimated through a majority rule consensus of the remaining

    samples. For this analysis we used only one individual from each

    populations using the Yule process as a speciation prior.

    Analyses based on a multi-locus coalescent approach enable a

    more accurate estimate of divergence processes, since gene diver-

    gence times normally occur before true speciation events

    (Maddison and Knowles, 2006; Carstens and Knowles, 2007;

    Edwards et al., 2007; Heled and Drummond, 2010; McCormack

    et al., 2010). Thus, during the species tree analysis, we set for the

    mtDNA dataset (cyt  b  and ND2 genes) the mutation rate of 2.1%substitutions per million years (Weir and Schluter, 2008), whereas

    those of the nuclear loci (Bf5/Bf7 and Musk4) were estimated in

    comparison to that of the mtDNA under a Yule process speciation

    prior.

     2.4. Population genetics, historical demography, and gene flow

    The recognized groups for the population genetics analyses

    were the reciprocally monophyletic populations revealed by the

    BI, ML, and species tree analyses (Fig. 1a and e). Even though pop-

    ulations of the Belém and Pernambuco areas of endemism did not

    show significant genetic differentiation and have not acquired reci-

    procal monophyly, they were recognized as independent popula-

    tions in the population genetics and network analyses, since they

    are currently separated by a long stretch of non-forest environ-

    ments (Cerrado and  Caatinga,  Ab’Saber, 1977), without possibility

    of current gene flow.We analyzed sequences of the three sampled loci (mtDNA, Bf5/

    Bf7 and Musk4) and populations with a minimum sampling of at

    least five individuals. The nuclear loci were used with both gametic

    phases, except females for Musk4. To estimate changes in historical

    demography through non-neutral models of evolution, we per-

    formed the Tajima’s  D   (Tajima, 1989), Fu’s  Fs   (Fu, 1997), and  R2

    Ramos-Onsins and Rozas (2002) tests. Significance was evaluated

    through 10,000 coalescent simulations. We also calculated nucleo-

    tide diversity (p) and haplotype diversity (h) for all populations. Allanalyses were performed in DNASP v.4.10.9 (Rozas et al., 2003).

    The dynamics of effective population sizes through time of the

    recognized populations of   T. aethiops   was estimated throughExtended Bayesian Skyline Plots (EBSP;   Heled and Drummond,

    2008), using a linear model in BEAST 1.6.1 (Drummond andRambaut, 2007). EBSP estimate changes in effective population

    sizes over time under a multi-locus approach by using the times

    of coalescent events among gene trees. We used the rate of 2.1%

    of nucleotide substitutions per million years (Weir and Schluter,

    2008) for the mtDNA with the relaxed clock and uncorrelated log-

    normal priors. The evolutionary rates of the nuclear loci (Bf5/Bf7

    and Musk4) were estimated in comparison to that the mtDNA.

    The MCMC parameters were the same used in the species tree

    analysis. The evolutionary models for each locus of each popula-

    tion were estimated in JMODELTEST 0.1.1 (Posada, 2008).

    To corroborate the obtained results in the phylogenetic analyses

    and test for population structuring across recognized geographical

    barriers (rivers and non-forest environments) we performed an

    analysis of molecular variance (AMOVA) using the mtDNA datasetin ARLEQUIN 3.1 (Excoffier et al., 2006).

    We applied the isolation with migration model (Hey and

    Nielsen, 2004) implemented in the IMa program (Hey and

    Nielsen, 2007) to estimate whether gene flow has occurred among

    those parapatric populations potentially in contact (Fig. 1b). For

    each pair of populations being compared, IMa estimates values of 

    H = 4N l  (where  l   is the mutation rate per year per marker) forpopulations 1 and 2, and their ancestor, as well as the   t   time

    (where t  = t l,  the time of population splitting at  t  generations inthe past) in which they diverged in the presence of gene flow(where m  = m/l, given in the coalescent). We structured the anal-yses assuming as populations the reciprocal monophyletic lineages

    reveled in the phylogenetic and species tree analyses. IMa esti-

    mates compared populations potentially in contact, without an

    apparent physical barrier separating them and parapatric popula-

    tions occurring on opposite banks of major Amazonian tributaries

    (see Fig. 1b). Our main objective with IMa analyzes was to test if 

    the observed absence of phylogenetic signal in the nuclear markers

    (see Fig. 1c) is better explained by incomplete lineage sorting as a

    result of recent population splitting (m = 0) or by the presence of 

    gene flow (m > 0). We assumed the Hasegawa–Kishino–Yano(HKY) model of evolution for all analyses and markers. Initial runs

    in MCMC mode were performed to establish the best priors for

    effective population sizes, divergence times, and migration param-

    eters for each pair of populations being compared. After prior

    selection, three independent final runs in MCMC mode were per-

    formed changing the seed number to observe the convergence of 

    parameter values. Every run was performed using 1,000,000 gener-

    ations of burn-in, sampling 100,000 trees in 5,000,000 generations

    with 20 chains with a geometric heating. We performed a likeli-

    hood ratio test (LRT,  Nielsen and Wakeley, 2001), using the Load-

    trees mode of IMa, sampling 100,000 generations and the results

    of the final MCMC runs, to reject or accept scenarios with migra-

    tion different from 0. If the LRT rejected a scenario of  m  > 0, weassumed that the absence of phylogenetic signal in the nuclear

    markers was due to incomplete lineage sorting. On the other hand,

    if the LRT rejected a scenario with  m   = 0, we performed another

    MCMC run fixing   m = 0 to evaluate whether a model withoutmigration adjusted better to the observed data than the full model

    with migration. For these comparison we applied the Akaike infor-

    mation Criterion (AIC), where the model that minimized

    AIC = 2[log (L) d] was chosen as the best, where d is the numberof parameters (Nielsen and Wakeley, 2001). If the model with

    migration had a better fit to the data, we assumed that the phylog-

    eographic pattern may have been affected by gene flow. To convert

    the model parameter estimations of  m  in demographic parameterswe applied the rate of 2.1% of nucleotide substitutions per million

    years based on the mtDNA calibration adopted (Weir and Schluter,

    2008), and a generation time of 1 year.

    3. Results

     3.1. Description of DNA sequences

    Results of Tajima’s   D   was not significant for the mtDNA and

    Musk4 (P  > 0.05), but did reject neutrality significantly for Bf5/Bf7(P  < 0.05; Table 3), suggesting that this locus does not evolve in aclock-like fashion. Wedid notdetect signsof saturation forthe faster

    evolving mitochondrial genes (cyt b and ND2). The phi’stest did notrejectthe null hypothesis of a dataset without recombination forthe

    nuclear genes (Bf5, Bf7 and Musk4; P  > 0.1). Were identified 57 dis-tinct haplotypes for cyt b   (992pb, n = 82), 38 for the ND2 (1041pb,

    n = 72), 42 for Bf5 (532pb, n = 118), 50 for Bf7 (963pb, n = 98), and30 for Musk4(565pb, n = 97; Table 3). All sequences weredeposited

    in Gen Bank under the accession numbers KF685940-KF686022

    (cyt   b), KF664029-KF664102 (ND2), KF686023-KF686081 (Bf5),KF686141-KF686186 (Bf7) and KF686081-KF686140 (Musk4).

    G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110   101

    http://-/?-

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    8/16

     3.2. Area and phylogenetic relationships

    BI and ML analyses based on the mtDNA and the concatenated

    dataset (total of four analyses; Fig. 1a) recovered identical topolo-

    gies showing with high statistical support the reciprocal mono-

    phyly of 10 populations of  T. aethiops, with T. aroyae  grouping asthe sister species. Two main clades were recovered in  T. aethiops,splitting apparently parapatric and attitudinally segregated

    Andean foothill from lowland Amazonian populations, which

    diverge by about 5% (uncorrected   p   genetic distance-value;Table 4). A subsequent major split involved lowland populations

    separated by the Tapajós River in south-central Amazonia, which

    diverge by roughly 4% and are hereafter referred to as the eastern

    and western clades (populations found to the east and west of the

    Tapajós River), respectively. The eastern clade grouped reciprocally

    monophyletic populations endemic to the Tapajós, Xingu, and

    Belém Amazonian areas of endemism (Silva et al., 2005), as well

    as the isolated Pernambuco population in the northern sector of 

    the Atlantic Forest biome in northeastern Brazil (Fig. 1a and b).

    The latter population was nested within the Belém population

    despite these two being located in different biomes separated by

    more than 1000 km of inhospitable habitat to T. aethiops (extensivenon-forest vegetation). The western clade grouped populations

    from the Madeira, Inambari, Napo, Imeri, and Guiana areas of ende-

    mism (Silva et al., 2005), but only the Napo/Imeri (west of the

    Negro River) and Guiana (east of the Negro River to west Branco

    River) areas were occupied by reciprocally monophyletic popula-

    tions (Fig. 1a and b). Although we did not sample populations from

    the Napo area of endemism (Silva et al., 2005), we tentatively

    include them in the Imeri population based on their morphological

    similarity and placement in the same subspecies   polionotus(Zimmer and Isler, 2003). In contrast, the Inambari and Rondônia

     Table 3

    Diversity and demographic parameters of lineages of the polytypic  T. aethiops  for which at least five individuals were sequenced for mitochondrial (cyt  b  and ND2) and nuclear

    (Bf5/Bf7, and musk4) genes. Lin. = lineages/clades; N = number of phased alleles; n H  = number of haplotypes;  H  = haplotype diversity; p = nucleotide diversity;  D = Tajima’s  Dtest (Tajima,1989);  F s = Fu’s Fs (Fu, 1997);   R2 = Ramos-Onsins, Rozas test (2002). Significance levels for Tajima’s  D  an d   R2 =  *P  < 0.05;   **P  < 0.01. Significance levels for Fu’s

    F s =  *P  < 0.02.

    Lin. Gene   N    n  H H    p   D F s   R2

    T. a. aethiops   mtDNA 72 64 0.996 0.0261 0.3502

    Bf5/Bf7 118 42 0.935 0.0054   1.68371*

    musk4 97 30 0.921 0.0051   1.37883

    Tapajós mtDNA 6 6 1.000 0.0018   0.4155   1.013 0.1446

    Bf5/Bf7 6 5 0.933 0.0061 0.7612   1.009 0.211

    musk4 5 3 0.700 0.0023   0.1747 0.0607 0.2848

    Xingu mtDNA 5 4 0.900 0.0014   0.1909   0.4448 0.2186

    Bf5/Bf7 6 4 0.867 0.0083   0.1057 1.0749 0.2073

    musk4 5 3 0.700 0.0033 0.2735 0.6436 0.1936

    Pernambuco mtDNA 6 4 0.800 0.0006   0.4474   1.4544 0.1805

    Bf5/Bf7 8 3 0.714 0.0037 0.5044 1.7263 0.2143

    musk4 6 3 0.600 0.0032 0.5617 0.957 0.2206

    Belém mtDNA 5 5 1.000 0.0042   0.4208   0.5418 0.1597

    Bf5/Bf7 12 6 0.864 0.0028 0.7877   1.8982 0.2027

    musk4 8 3 0.714 0.0024 0.9657 0.8748 0.2381

    Eastern IInambari mtDNA 6 6 1.000 0.0021   0.6915   2.4843 0.1446

    Bf5/Bf7 6 6 1.000 0.0077   0.9902   2.4199 0.1023**

    musk4 6 5 0.933 0.0044 1.0484   1.5652 0.2386

    Western Inambari mtDNA 15 13 0.981 0.0021   1.5705* 7.6834* 0.0700*

    Bf5/Bf7 32 14 0.897 0.0034   1.6007* 8.3201* 0.0582**

    musk4 25 11 0.847 0.0031   0.6854   5.3341* 0.0958

    Northern Rondônia mtDNA 7 7 1.000 0.0021   0.3606   3.4573* 0.1267*

    Bf5/Bf7 14 8 0.901 0.0040   0.768   2.8844 0.1174

    musk4 11 9 0.945 0.0045   0.9471   5.1712* 0.1247

    Southern Rondônia mtDNA 15 12 0.971 0.0047   0.4438   2.11639 0.1208

    Bf5/Bf7 24 6 0.757 0.0023   1.2115   1.1177 0.1042

    musk4 20 7 0.800 0.0036   0.5208   0.9804 0.1166

     Table 4

    Average pairwise uncorrected p-distances (mtDNA) between and within major clades of the polytypic  Thamnophilus aethiops and outgroups recovered by Bayesian and Maximum

    Likelihood phylogenies.

    Clade (average pairwise p-distances within clades) 1 2 3 4 5 6 7 8 9 10 11

    1 – Western Negro (0.1)

    2 – Eastern Negro (0) 0.9

    3 – Western Inambari (0.2) 1.6 1.4

    4 – Southern Rondônia (0.6) 2.2 2.1 2.7

    5 – Eastern Inambari (0.2) 2.2 2.3 2.9 1.7

    6 – Northern Rondônia (0.2) 2.7 2.8 3.1 3.0 3.4

    7 – Tapajós (0.1) 3.6 3.7 4.0 3.8 3.9 3.9

    8 – Xingu (0.2) 3.4 3.4 3.6 3.6 3.7 3.8 1.9

    9 – Belém (0.4) 3.1 3.1 3.5 3.2 3.2 3.4 1.9 0.7

    10 – Pernambuco (0.1) 3.1 3.2 3.5 3.3 3.4 3.3 1.8 0.6 0.3

    11 –  T. a. aethiops   5.0 5.1 5.7 5.2 5.3 6.0 4.3 4.8 4.5 4.7

    12 –  T. aroyae   5.2 5.5 6.2 5.2 5.5 5.5 5.2 5.8 5.5 5.6 6.0

    102   G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    9/16

    areas of endemism were inhabited by four distinct non-sister pop-

    ulations with the following distributions: (1) northern portion of 

    the Rondônia area of endemism in the Tapajós–Madeira interfluve,

    probably limited to the south by the Aripuanã River; (2) southern

    part of the Rondônia area of endemism, probably limited to the

    north by the Aripuanã River, crossing the Tapajós–Madeira inter-

    fluve eastwards into the headwaters of the Tapajós and Xingu Riv-

    ers, and hence reaching the southernmost parts of the Tapajós andXingu areas of endemism; (3) eastern Inambari area of endemism

    in the lower–central portion of the Purus–Madeira interfluve; and

    (4) western Inambari area of endemism west of the Purus River

    and south of the Marañon/Huallaga/Ucayali/Solimões River in

    Peru, eastward reaching the headwaters of the Purus–Madeira

    interfluve into northeastern Bolivia (Pando and La Paz Depart-

    ments). Uncorrected genetic  p-distances among populations of  T.aethiops ranged from 0.6% (between Belém/Pernambuco and Xingupopulations) to 6.0% between Andean foothill and northern Rondô-

    nia area of endemism populations, with the average  p-distancesbetween reciprocally monophyletic sisters populations around

    1.6%.

    The coallescent species tree recovered an identical topology to

    those of BI and ML analyses based on the concatenated dataset

    (Fig. 1e). The species tree also recovered with high statistical sup-

    port the monophyly of the polytypic  T. aethiops, as well as that of the Amazonian populations, the eastern and western clades, and

    the internal relationships within them. Only the sister relationship

    between the Western Inambari and the Northern Solimões River

    populations was not statistically supported (posterior probability

    less than 0.95; Fig. 1e).

    The individual topologies of each nuclear gene (Bf5, Bf7 and

    Musk4, results not shown) do not support any particular well-

    resolved relationship among individuals of   T. aethiops, except forMusk4, which supported the monophyly of lowland Amazonian

    and Atlantic Forest populations.

     3.3. Species delimitation

    We obtained equal results among all six BP&P analyses, using

    both algorithms and the three different combinations of  s   and  hvalues, assuming a cutoff of 95% of posterior probability. Species

    delimitations results were robust to the choice of prior distribu-

    tions, supporting most of the reciprocally monophyletic popula-

    tions recovered by the BI, ML, and species tree estimates as

    distinct species, except for the Belém/Pernambuco and Eastern/

    Western Negro sister population pairs (Fig. 1d). The purported spe-

    ciation event between Belém and Pernambuco populations had

    higher speciation probabilities when we used priors representing

    large ancestral populations and deep divergences (PP = 0.90) or

    shallow divergences (PP = 0.89), but very low values with priors

    representing small ancestral populations and shallow divergences

    (PP = 0.10), using algorithm 0 with the fine-tuning parameter setto 5. The split between Eastern and Western Negro populations

    had very low speciation probabilities in all analyses (PP 6 0.20).

     3.4. Molecular dating 

    According to the molecular dating based on the concatenated

    mtDNA dataset (Fig. 1e), the split between T. aethiops and T. aroyaetook place during the Pliocene between 4.07 and 2.85 Mya (mean

    3.45 Mya). The divergence between the Andean (T. a. aethiops) andthe ancestor of the Amazonian and Atlantic Forest populations took

    place between 3.54 and 2.44 Mya (mean 2.97 Mya).

    The earliest split within Amazonian populations occurred in the

    Late Pliocene–Early Pleistocene (2.70–1.86 Mya, mean 2.27 Mya),

    with separation of the eastern and western clades across theTapajós River. The first split in the western clade separated

    populations from the northern part of the Rondônia area of ende-

    mism (east of Aripuanã River) from the remaining ones found to

    the west of this river and across the Madeira, Solimões, and Negro

    rivers, during the Early–Middle Pleistocene (1.84–1.19 Mya, mean

    1.50 Mya). Other cladogenetic events in the western clade

    stretched from the Middle to Late Pleistocene between 1.4–0.29

    Mya (mean 0.84 Mya), with one of them (involving southern

    Rondônia + eastern Inambari versus western Inambari + northernSolimões populations) coinciding to some extent with the modern

    course of the Purus River (Fig. 1b and e). The splits in the eastern

    clade occurred between the Middle (1.27–0.66 Mya, mean 0.95

    Mya) and Late Pleistocene (0.49–0.15 Mya, mean 0.31 Mya) and

    separated populations of the Tapajós, Xingu, and Belém/Pernam-

    buco areas of endemism across the Xingu and Tocantins rivers.

    The splitting dates estimated by the multi-locus coalescent

    approach (species tree) were in general more recent than those

    obtained by the single locus mtDNA dataset, particularly for those

    younger than 1.0 Mya. However, both coalescent and concatenated

    mtDNA timing estimates overlapped to various degrees in their

    respective 95% confidence intervals, and thus cannot be considered

    statistically different.

     3.5. Population genetics and historical demography

    Haplotype networks recovered for cyt  b  and ND2 mirrored thephylogenetic analyses based on BI and ML (Fig. 1a and c). For cyt

    b,   populations with the highest number of sequenced individuals

    differed dramatically in structure: the western Inambari popula-

    tion presented a dominant haplotype shared by six individuals

    and 12 other haplotypes represented by one or two individuals,

    generally differing from the commonest haplotype by a single

    mutation. On the other hand, in the southern Rondônia population,

    10 different haplotypes distributed in three groups separated by

    two to eight mutations were recovered, with only three haplotypes

    present in more than one individual (Fig. 1c). Networks of the

    nuclear makers recovered the absence of phylogeographic struc-

    ture and haplotype sharing among the recognized populations(Fig. 1c).

    Haplotype diversity was generally high and similar among the

    sampled Amazonian populations (Table 3). Estimated nucleotide

    diversity varied widely among populations depending on the locus

    considered, while within a single locus values obtained for the dif-

    ferent populations were similar (Table 3), suggesting an overall

    demographic stability. The Tajima’s D test (1989) showed negative

    and significant values for the mtDNA and Bf5/Bf7 loci in the wes-

    tern Inambari population, consistent with demographic fluctua-

    tions (Table 3). The Fu’s Fs (1997) showed significant and

    negative values for the mtDNA and musk4 loci in the northern

    Rondônia population and in all three sampled loci in the western

    Inambari, also supporting demographic instability. The  R2 (2002)

    test results showed low and significant values for the same popu-lations as the Fu’s Fs (1997), suggesting a similar demographic sce-

    nario regardless of the test considered.

    The EBSP’s (Fig. 2) were in part congruent with the neutrality

    tests, inferring tendencies of recent demographic expansions in

    both populations of the Inambari and the northern Rondônia areas

    of endemism during the Late Pleistocene and Holocene. In contrast,

    EBSP’s favored scenarios of demographic stability in the Belém,

    Pernambuco, Xingu, Tapajós, and southern Rondônia populations

    (Fig. 2).

    An analysis of molecular variation (AMOVA; Table 5) confirmed

    that most of the genetic variation in the Amazonian and Atlantic

    Forest populations of   T. aethiops   is partitioned among the geo-

    graphically structured clades recovered by the phylogenetic analy-

    ses, and not within these groups. Similarly, these resultscorroborate that the large tributaries of the Amazon River are at

    G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110   103

    http://-/?-http://-/?-

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    10/16

    least contemporary barriers to gene flow between populations

    from opposite banks.

     3.6. Gene flow among populations

    IMa analyses favored scenarios of very low levels or absence of 

    gene flow among all populations of  T. aethiops  potentially in con-

    tact (Table 6). Migration events were not rejected by LRT tests

    among most pairwise comparisons, with models predicting migra-tion having better AIC scores than models predicting no gene flow

    between the following clades: Xingu/Tapajós, Northern Rondônia/

    Southern Rondônia, Northern Rondônia/Eastern Inambari, South-

    ern Rondônia/Western Inambari, and Eastern Inambari/Western

    Inambari. In all these instances, the estimated migration rate was

    very low in both directions (with confidence intervals’ lower bonds

    reaching zero except in one comparison), and involved non-sister

    populations (Fig. 1,  Table 6). The highest migration rate detected

    and the only one not including a lower bond likelihood of zero,

    involved the Eastern Inambari and Western Inambari populations(Maximum likelihood = 0.0015; 90% HPD interval = 0.001–0.004).

    Fig. 2.   Demographic histories of major clades of  Thamnophilus aethiops   for which at least five individuals were sampled, inferred through Extended Bayesian Skyline Plots

    based on mtDNA (cyt b  and ND2), Bf5/Bf7, and Musk4 sequences. Black solid lines represent median values, while the dashed lines corresponds to 95% confidence intervals.

    The X axis corresponds to time in million years before present. The Y axis represents  N es   in a logarithmic scale.

    104   G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    11/16

    4. Discussion

    4.1. Species limits and taxonomy

    Although our analyses support the monophyly of the polytypic

    T. aethiops, of the 10 reciprocally monophyletic populations recov-ered only two are entirely consistent with subspecies limits cur-

    rently recognized in this group: nominate   aethiops   from thefoothills of the Andes and   T. a. injunctus, which corresponds tothe eastern Inambari population (Zimmer and Isler, 2003; Fig. 1a,

    d and e). In fact, high statistical support for key nodes in the

    multi-locus coalescent species tree and high speciation probabili-

    ties recovered by the species delimitation analysis (BP&P), support

    the recognition of nine reciprocally monophyletic lineages fitting

    the definition of evolutionary species (see   Fujita et al., 2012;Fig. 1d) in T. aethiops.

    These lineages are as follows (taxon names are assigned torecovered populations based on type locality and nomenclatural

    priority following  Peters 1951   and  Zimmer and Isler 2003): (1)

    Thamnophilus aethiops   Sclater, 1858 (foothills of the Andes); (2)Thamnophilus atriceps Todd, 1927 (Tapajós population, distributedthroughout most of the Tapajós area of endemism); (3) Thamnophi-lus incertus   Pelzeln, 1869 (Belém and Pernambuco populations;

    given their lack of reciprocal monophyly and low speciation prob-

    abilities, the former Thamnophilus aethiops incertus from the Belémarea of endemism and   Thamnophilus aethiops distans  Pinto 1954from the Atlantic Forest should be regarded as subspecies of a sin-

    gle species, T. incertus, the name with nomenclatural precedence);(4)   Thamnophilus punctuliger   Pelzeln 1869 (northern Rondôniapopulation distributed east of the Aripuanã River in the Madeira–

    Tapajós interfluve); (5)   Thamnophilus injunctus   Zimmer 1933(eastern Inambari population, distributed through most of the

     Table 5

    Results of an Analysis of Molecular Variation (AMOVA) based exclusively on the mtDNA among major clades of the polytypic  Thamnophilus aethiops   for which at least five

    individuals were sampled, separated by potential modern barriers (rivers and non-forest habitats). Results regarded as significant if  P  < 0.001.

    Populations Potential barrier Among p opulations o n opposite s ides o f  

    the barrier

    Within the population in the same side

    of the barrier

    Ust   p-

    Value

    Pernambuco/Belém   Caatinga/Cerrado (non-forest

    landscapes)

    18.77 81.23 0.18 0.026

    Belém/Xingu Tocantins R. 60.42 39.58 0.60 0.004

    Xingu/Tapajós Xingu R. 91.08 8.92 0.91 0.002Xingu/Southern Rondônia Unknown 84.78 15.22 0.84 0.000

    Tapajós/Northern Rondônia Lower Tapajós R. 71.25 28.75 0.71 0.000

    Tapajós/Southern Rondônia Teles Pires R. (part) 46.49 53.51 0.46 0.000

    Northern/Southern Rondônia Aripuanã R. (part) 86.05 13.95 0.86 0.000

    Northern Rondônia/Eastern

    Inambari

    Lower Madeira R. 93.39 6.61 0.93 0.000

    Southern Rondônia/Eastern

    Inambari

    Middle Madeira R. 73.82 26.18 0.73 0.000

    Southern Rondônia/Western

    Inambari

    Upper Madeira R. 84.84 15.16 0.84 0.000

    Eastern/Western Inambari Purus R. (part) 92.26 7.74 0.92 0.000

     Table 6

    Summary of IMa results of migration estimates between parapatric population pairs of  T. aethiops  potentially in contact (see Fig. 1b). Posterior distribution values of migration

    were converted to demographic units, where the values of  m1 and  m2 represent the average number of migrations per 1000 generations per gene copy on the coalescent

    (m1 = rate of migration from population 2 to population 1;  m2 = rate of migration from population 1 to population 2). Confidence intervals represent  HPD90. All maximum values

    for migration reached likelihood zero, indicating the posterior distribution ends.

    Pairwise analyses   m1   m2 AIC

    Belém (5)/Xingu (5) 0 (0–0.0011) 0.0027 (0–0.0063) 9569.71

    Forced to be zero Forced to be zero   9548.03

    Xingu (5)/Tapajós (6) 0 (0–0.0033) 0 (0–0.003)   9586.18

    Forced to be zero Forced to be zero 9613.29

    Xingu (5)/Southern Rondônia (15)* 0 0 –

    Tapajós (6)/Northern Rondônia (7)* 0 0 –

    Tapajós (6)/Southern Rondônia (15)* 0 0 –

    Northern Rondônia (7)/Southern Rondônia (15) 0 (0–0.0012) 0 (0–0.0013)   10366.19

    Forced to be zero Forced to be zero 10378.84

    Northern Rondônia (7)/Eastern Inambari (6) 0 (0–0.0013) 0 (0–0.002)   9996.38Forced to be zero Forced to be zero 10012.88

    Southern Rondônia (15)/Eastern Inambari (6) 0 (0–0.0016) 0.0013 (0–0.0032) 9920,21

    Forced to be zero Forced to be zero   9898.74

    Southern Rondônia (15)/Western Inambari (9) 0 (0–0.0019) 0 (0–0.0012)   10963.26

    Forced to be zero Forced to be zero 10997.83

    Eastern Inambari (6)/Western Inambari (9) 0 (0–0.0032) 0.0015 (0.001–0.004)   10369.91

    Forced to be zero Forced to be zero 10395.45

    Western Inambari (9)/Northern Solimões (7)* 0 0 –

    Asterisks indicate pairwise analyses where the likelihood ratio test rejects a scenario with m > 0. For pairwise comparisons where the likelihood ratio test supports a scenario

    with m > 0, models with and without migration were contrasted with the Akaike information criterion (AIC). The model that minimized AIC was chosen as the best scenario

    (in bold). Numbers in parenthesis next to population names denote the number of individuals of the respective population used in the comparison. See  Table 1 for detailed

    specimen and locality information.

    G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110   105

    http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    12/16

    Madeira–Purus interfluve); (6)   Thamnophilus juruanus   Iheringi1904 (western Inambari population, distributed through most of 

    the Inambari area of endemism west of the Purus River; the taxon

    name  Thamnophilus aethiops kapouni  Seilern 1913 also applies to

    this population, but it has no nomenclatural priority, and therefore

    should be synonymized with juruanus); (7)  Thamnophilus poliono-

    tus  Pelzeln 1869 (western and eastern Negro populations distrib-

    uted north of the Amazon in the Branco–Solimões interfluve; asexplained above the taxon   Thamnophilus aethiops wetmorei   deSchauensee, 1945 is tentatively included in this lineage); and,

    finally, two additional unnamed lineages (no known taxon

    described previously from their ranges) corresponding respectively

    to most of the Xingu area of endemism (assigned currently to  T.atriceps, whose type locality is in the Tapajós area of endemism)and the southern part of the Rondônia, Tapajós, and Xingu areas

    of endemism (assigned currently to   T. punctuliger , whose typelocality lies in the northern part of the Rondônia area of 

    endemism).

    By occupying mostly shaded environments in the understory of 

    tropical forests, the evolution of plumage color states traditionally

    used in subspecies delimitations in  T. aethiops probably proceededat a slower pace than at least one of the sequenced loci (mtDNA).

    This process can generate the formation of morphologically cryptic

    and sometimes polymorphic lineages, resulting in the taxonomic

    inconsistencies cited above (Baker et al., 1995; Irwin et al., 2001;

    Fernandes et al., 2012). In contrast, vocal characters are thought

    to be directly involved in the reproductive isolation of forest inte-

    rior suboscine lineages (Carneiro et al., 2012). We do not report on

    the vocal variation within the polytypic  T. aethiops   in this paper,but will present in a separate publication a detailed taxonomic

    overhaul of the  T. aethiops   complex based on a multi character

    approach, where the unnamed evolutionary lineages mentioned

    above will be described.

    The only sister populations of  T. aethiops for which low specia-tion probabilities were recovered are the parapatric eastern/wes-

    tern Negro populations (T. polionotus) and the allopatric Belém/

    Pernambuco populations (T. i. incertus   and   T. i. distans). Eventhough the species tree recovered the former pair as distinct spe-

    cies, the species delimitation analyses favored a scenario of very

    shallow divergence between them, not supporting their status as

    distinct evolutionary species (Fig. 1d and e). Indeed, the lack of 

    reciprocal monophyly in the mtDNA between the Belém (T. i. incer-tus) and Pernambuco (T. i. distans) populations supports a very

    recent divergence between them, despite their current separation

    by over 1000 km.

    The overall lack or very low rate of gene flow estimated among

    most parapatric populations of  T. aethiops   are also indicative of 

    their status as independent species (Table 6). Interestingly, when-

    ever lack of gene flow (i.e., zero migration rates) could be statisti-

    cally rejected between any population pair potentially in contact, it

    always involved non-sister lineages, implying in secondary contactrather than primary intergradation (Table 6). Hence, it is possible

    that these very low rates of gene flow are explained by a significant

    degree of reproductive isolation already acquired by these popula-

    tions that have diverged in allopatry for a relatively long time

    before coming in contact. Estimates of gene flow between non-sis-

    ter populations or populations that exchange migrants with a third

    unsampled population go against a main assumption of the IMa

    model and can generate several biases in the estimation of the

    main parameters, although they have little effect when low levels

    of gene flow are detected (see   Strasburg and Rieseberg 2010), as

    documented herein. IMa analyses supported absolute no gene flow

    among all populations of the two main Amazonian clades of   T.aethiops separated by the Tapajós River (eastern clade: Xingu and

    Tapajós populations; and western clade: northern Rondônia andsouthern Rondônia populations), which likely come in contact

    across a wide area in the headwaters of the Tapajós, Xingu, and

    Tocantins rivers (Fig. 1b,   Table 6). Despite sampling limitations,

    this study provides the best evidence available so far that interflu-

    vial populations of  terra-firme avian lineages in contact across the

    headwaters of southern Amazonian tributaries can retain their

    genetic diagnosabilities, probably as a result of an advanced degree

    of reproductive isolation acquired in allopatry before coming in

    contact in more recent times, as suggested by several other studiesbased solely on mtDNA datasets (Carneiro et al., 2012; Batista

    et al., 2013; Rodrigues et al., 2013). In contrast, gene flow was pres-

    ent, yet at a very small rate, between the non-sister Tapajós and

    Xingu populations belonging to the same eastern clade, as well

    as between the southern Rondônia and western Inambari popula-

    tions of the western clade across the upper Madeira River ( Fig. 1,

    Table 6). Therefore, comparative levels of gene flow across the

    headwaters of major Amazonian tributaries are likely to be more

    correlated with the phylogenetic distance of the lineages/popula-

    tions in contact, rather than the smaller barrier effect of rivers in

    these areas (Haffer 1992), which result in a more random pattern

    of gene flow.

    4.2. The diversification of the Thamnophilus aethiops complex in Cis- Andean South America

    Populations of  T. aethiops distributed in the Amazon and Atlan-tic Forests are in a clade with several basal Andean taxa typically

    occurring above 1000 m (T. unicolor, T. aroyae, and T. a. aethiops),

    supporting an ancestor of Andean origin for these lowland forest

    populations (see also   Brumfield and Edwards, 2007). Diversifica-

    tion events along Andean altitudinal gradients have been proposed

    for some avian lineages (Bates and Zink, 1994; Weir, 2006; Isler

    et al., 2012; d’Horta et al., 2013), and are probably related to allo-

    patric speciation during altitudinal shifts in habitats due to cli-

    matic oscillations (Graham et al., 2004). Under this diversification

    model, the dispersal of an Andean ancestor into the Amazonian

    lowlands would only be possible after the draining of the lacus-

    trine system of the Solimões Basin in western Amazonia (the socalled Pebas and Amazonas lakes), which led to the formation of 

    the modern Amazon drainage (Hoorn et al., 2010) and the estab-

    lishment of upland   terra-firme   forests in this region. Paleogeo-graphic hypotheses differ in the time frame when these events

    occurred (Campbell et al., 2006; Figueiredo et al., 2009;

    Latrubesse et al., 2010), even though they coincide in postulating

    a favorable environment for the establishment of upland   terra- firme  forests in western Amazonia between the late Pliocene andearly Pleistocene, when the diversification of the  T. aethiops  com-plex in Amazonia began (Fig. 1e).

    The genetically poorly-differentiated yet isolated Atlantic Forest

    population (T. i. distans) may have originated from a recent dis-persal event of the Belém population (T. i. incertus) or through iso-

    lation caused by climatic fluctuations leading to the disruption of the forest corridor linking eastern Amazonia with the northeastern

    Atlantic Forest during the Late Quaternary (Batalha-Filho et al.,

    2013). In contrast, other Amazonian and Atlantic Forest avian sister

    lineages exhibited reciprocal monophyly and high levels of genetic

    differentiation estimated to date back to a time frame stretching

    from the Late Pliocene to the Early Pleistocene (Aleixo, 2002;

    Cabanne et al., 2008; Weir and Price, 2011). These different phylog-

    eographic patterns are consistent with multiple and cyclical con-

    nection and isolation episodes between eastern Amazonia and

    the northern sector of the Atlantic Forest until at least the Late

    Quaternary (Batalha-Filho et al., 2013).

    The data presented herein support that distinct events of vicar-

    iance related to river formation and reorganization mostly during

    the Pleistocene have influenced the diversification of the T. aethiopscomplex, providing another body of evidence that different

    106   G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    13/16

    modern interfluves in Amazonia harbor distinct evolutionary lin-

    eages, with little or no gene flow among them (Table 6). In the  T.

    aethiops complex, all splits involving the recognized lineages coin-cide with at least part of the course of a major Amazonian river,

    indicating that at some point in time these rivers might have acted

    as primary diversification barriers. Interestingly, the absence of a

    physical barrier separating the eastern and western Inambari

    clades where they meet in the middle Purus–Madeira interfluveis associated with the highest gene flow rate recovered herein

    among lineages of the   T. aethiops   complex, yet still very low(0.0015; Table 6).

    The phylogeographic data presented herein is consistent with at

    least two independent splits across different portions of the mod-

    ern course of the Madeira River as also suggested for several other

    avian lineages (Aleixo 2004; Patané et al., 2009; Fernandes et al.,

    2013; Sousa-Neves et al., 2013). Latrubesse (2002) postulated that

    the Madeira paleodrainage was broader and more complex than

    today until the late Pleistocene, reporting the presence of two

    mega-fans related to the Aripuanã and Jiparaná rivers, both of 

    which currently flow into the Madeira River. Thus, it is possible

    that the initial isolation of the northern Rondônia lineage in the

    western clade is related to a former larger barrier (Aripuanã

    mega-fan), which decreased in size due to a drainage-capture

    event (Wilkinson et al., 2010), promoting the secondary contact

    between different non-sister lineages that diverged in allopatry,

    as postulated for other   terra-firme   avian lineages (Fernandes2013). The absence or very low levels of gene flow detected among

    populations of the T. aethiops  complex suggest a fickle scenario of migration across Amazonian rivers, which is probably best

    explained by an overall inability to cross open water, as demon-

    strated for other understory Neotropical avian lineages (Moore

    et al., 2008). Thus, multiple and independent splitting events along

    a single modern river course are more likely explained by the dis-

    ruption of river drainages, whereby ‘‘barrier effects’’ move across

    neighbouring interfluvia.

    Likewise, the estimated dates for the splits across the Solimões

    and Negro rivers long after the formation of these barriers, suggesta similar scenario of river lateral change or course capture. The for-

    mation of the Solimões River is closely related to the formation of 

    the transcontinental Amazonian drainage during the Late Plio-

    cene–Early Pleistocene (Campbell et al., 2006; Hoorn et al., 2010;

    Latrubesse et al., 2010). Thus, if these time estimates for the estab-

    lishment of the Solimões River are correct, the relatively recent

    split of   T. aethiops  populations across this river (0.68–0.26 Mya)is explained by vicariance following dispersal after the formation

    of the current Solimões River. As observed for clades separated

    by the Solimões, the formation of the Negro River (see   Almeida-

    Filho and Miranda 2007) probably predates the cladogenetic event

    separating the Imeri/Napo and western Guianan clades of T. polion-otus   (0.23–0.02 Mya). Unlike widely reported local instances of 

    river-channel migration due to sedimentation involving ‘‘white-water’’ rivers in western Amazonia (Salo et al., 1986; Patton and

    Silva, 1998), some shifts in Amazonian rivers courses take place

    over long periods of time and are probably related to drainage cap-

    ture events mediated by tectonics (Almeida-Filho and Miranda

    2007; Rossetti and Valeriano, 2007; Shephard et al., 2010). The

    growing body of phylogeographic evidence available now for Ama-

    zonian lineages suggest that river courses, particularly those of the

    Negro, Madeira, Tapajós and Tocantins, have been stable for long

    periods of time until changing part of their courses, and then

    remaining stable for an additional time (Ribas et al., 2012;

    d’Horta et al., 2013; Fernandes et al., 2012, 2013; Maldonado-

    Coelho et al., 2013; Sousa-Neves et al., 2013). Hence, a dynamic

    drainage landscape, over long periods of time, can act like a speci-

    ation pump rather than prevent diversification, as suggested previ-ously (Haffer 1993; Colwell 2000; Gascon et al., 2000).

    When compared to other upland   terra-firme   avian lineageswhich were also shown to have their diversification mediated by

    Amazonian rivers (Fernandes et al., 2012, 2013; Ribas et al.,

    2012; d’Horta et al., 2013; Sousa-Neves et al., 2013), the order

    and timing in which different rivers accounted for splits of lineages

    of the  T. aethiops   complex differ significantly from most of them,indicating that distinct lineages responded differently to the pro-

    cess of drainage evolution in Amazonia. Perhaps the most similarcases to that of the  T. aethiops  complex are those reported for the Xiphorhynchus spixii/elegans  (Dendrocolaptidae;   Aleixo 2004) and

    Phlegopsis nigromaculata   (Thamnophilidae;   Aleixo et al., 2009),whereby the Tapajós River also accounted for the first split in time.

    One hypothesis is that these differences in the relative order and

    timing in which different rivers account for splits throughout the

    Amazon are related to the geografic origins of the ancestral popu-

    lations of these widespread lineages (Aleixo and Rossetti 2007). In

    contrast to the  T. aethiops   complex,  X. spixii/elegans and Sclerurus

    mexicanus (d’Horta et al., 2013), whose sister groups in Amazoniainclude widespread Andes lineages, a different but pervasive phy-

    logeographic pattern in the Amazon is that where earliest splits

    usually involve lineages endemic to the Guianan shield and coin-

    cide either with the present course of the lower Amazon or Negro

    rivers (Carneiro et al., 2012; Batista et al., 2013; d’Horta et al.,

    2013; Rodrigues et al., 2013; Sousa-Neves et al., 2013). This alter-

    native pattern imply a Guianan shield origin for these lineages,

    whereas those reported for T. aethiops, P. nigromaculata, and X. spi- xii/elegans   (which are absent or nearly absent from the Guianan

    shield) support a Brazilian shield origin (see also   Aleixo and

    Rossetti 2007). Therefore, we hypothesize that the structuring of 

    the modern Amazon drainage and the drying out of the western

    Amazonian sedimentary basins from the late Pliocene to the late

    Pleistocene were keystone events to promote concomitantly: (1)

    an increase in the dispersal rates of both Brazilian and Guianan

    shield lineages across the more recent upland  terra-firme   forestsof the upper Amazonia (Solimões sedimentary basin); and (2) an

    increase in cladogenetic events involving upland  terra-firme forest

    lineages due to the latest cycle of drainage evolution throughoutAmazonia, which shaped the modern transcontinental river sys-

    tem. Under this scenario, depending on the geographical origins

    of any given lineage before this latest cycle of widespread drainage

    structuring (late Pliocene–Early Pleistocene), their response to riv-

    erine barriers can differ widely, as verified by the phylogeographic

    studies conducted so far.

    When climate change is evaluated as a possible driver of diver-

    sificationin the T. aethiops complex, neutrality tests and EBSP’s sup-port demographic expansions in lineages distributed in the

    Inambari and northern Rondônia areas of endemism since the last

    interglacial cycle (approximately 0.13 Mya) to the present (Fig. 2).

    Recent demographic expansions have been postulated for other lin-

    eages of birds in western Amazonia (Aleixo, 2004; Fernandes et al.,

    2012; Ribas et al., 2012), but in all cases there was no evidence thatpurported refuges were correlated with cladogenetic events in

    these lineages. The dates estimated for the origin of the main recip-

    rocally monophyletic lineages of the T. aethiops complex based on

    the multi-locus species tree precede the coalescent time of the

    locus used in EBSP’s analyzes as well as the estimated time of the

    demographic expansions. Thus, if climatic oscillations during the

    last 0.2 Mya affected populations of the  T. aethiops  complex, theyapparently were not strong enough to generate the formation of 

    new clades, hence not supporting the refuge hypothesis as an

    important cause of diversification in the   T. aethiops   complex(Haffer, 2001). Furthermore, differences in the historical population

    demography among   T. aethiops   complex lineages (Fig.  1c and 2),

    with some lineages exhibiting relatively stable demographic histo-

    ries, suggest that climate oscillations during the Late Pleistocenedid not impact uniformly all sectors of Amazonia. Even though

    G. Thom, A. Aleixo / Molecular Phylogenetics and Evolution 82 (2015) 95–110   107

  • 8/16/2019 Cryptic Speciation in the White Shouldered Antshrike Thamnophilus Aethiops Aves Thamnophilidae the Tale of a Tr…

    14/16

    the sparse sampling of some T. aethiops complex lineages may yieldunreliable results that underestimate the effect of climatic oscilla-

    tions on ancestral population sizes during the last glacial cycles,

    our data is in agreement with that recovered for several other

    linages of Amazonian organisms where no significant changes in

    historical population structures were found since the Last Glacial

    Maximum (Lessa et al., 2003; Fernandes et al., 2012; Ribas et al.,

    2012; Maldonado-Coelho et al., 2013). These results have ignitedan intense debate on the applicability of the refuge hypothesis in

    explaining patterns of diversification in Amazonia (Bush and de

    Oliveira, 2006), along with paleo-environmental uncertainties con-

    cerning the existence and location of refuges (Bush, 1994; van der

    Hammen and Hooghiemstra, 2000; Haffer, 2001).

     Acknowledgments

    We thank the curator and curatorial assistants of the Museum

    of Natural Science, Louisiana State University (LSUMZ), for allow-

    ing us to sequence tissues and study specimens under their care.

    We also grateful to the tenacious effort of many specimen collec-

    tors working for or together with the Museu Paraense Emílio Goe-

    ldi (MPEG), who along years of intense field work in the Amazon,

    amassed the tissues necessary for the analyzes contained in this

    paper, with permits provided by ICMBIO (Instituto Chico Mendes

    de Conservação da Biodiversidade). Field and laboratory work

    related to this paper were generously funded by the Brazilian

    Research Council (CNPq; grants # 310593/2009-3; ‘‘INCT em Bio-

    diversidade e Uso da Terra da Amazônia’’ # 574008/2008-0; #

    563236/2010-8; and # 471342/ 2011-4) and the Pará State Funding

    Agency (ICAAF 023/2011). Support to GT’s graduate research was

    provided by a CNPq Master’s fellowship (# 131000/2010-1). AA

    is supported by a CNPq research productivity fellowship (#

    310880/2012-2). We thank an anonymous referee as well as Mar-

    cos Maldonado-Coelho, Renato Caparroz, Luciano N. Naka, Camila

    Ribas, and Péricles S. Rego for reviewing earlier versions of this

    manuscript.

    References

    Ab’Saber, A.N., 1977. Os domínios morfoclimáticos da América do Sul. Primeiraaproximação. Geomorfologia 53, 1–23.

    Aleixo, A., 2002. Molecular systematics and the role of the ‘‘várzea’�