DIVERSIFICAÇÃO EVOLUTIVA EM ANFÍBIOS AMAZÔNICOS
Transcription
DIVERSIFICAÇÃO EVOLUTIVA EM ANFÍBIOS AMAZÔNICOS
INSTITUTO NACIONAL DE PESQUISAS DA AMAZÔNIA - INPA PROGRAMA DE PÓS-GRADUAÇÃO EM ECOLOGIA DIVERSIFICAÇÃO EVOLUTIVA EM ANFÍBIOS AMAZÔNICOS (Allobates, DENDROBATOIDEA): CARACTERES GENÉTICOS, MORFOLÓGICOS E COMPORTAMENTAIS IGOR LUIS KAEFER Manaus, Amazonas Outubro, 2012 IGOR LUIS KAEFER DIVERSIFICAÇÃO EVOLUTIVA EM ANFÍBIOS AMAZÔNICOS (Allobates, DENDROBATOIDEA): CARACTERES GENÉTICOS, MORFOLÓGICOS E COMPORTAMENTAIS ORIENTADORA: Dra. ALBERTINA PIMENTEL LIMA CO-ORIENTADORA: Dra. IZENI PIRES FARIAS Tese apresentada ao Instituto Nacional de Pesquisas da Amazônia como parte dos requisitos para obtenção do título de Doutor em Biologia (Ecologia). Manaus, Amazonas Outubro, 2012 ii BANCA EXAMINADORA DO TRABALHO ESCRITO Nome (instituição) Parecer Andrew Crawford (Smithsonian Tropical Research Institute) Não emitido Heike Pröhl (Stiftung Tierärztliche Hochschule Hannover) Aprovado Jeffrey Podos (University of Massachusetts) Aprovado Vanessa Verdade (Universidade Federal do ABC) Aprovado Walter Hödl (Universität Wien) Aprovado BANCA EXAMINADORA DA DEFESA PÚBLICA DA TESE Nome (instituição) Parecer Camila Ribas (Instituto Nacional de Pesquisas da Amazônia) Aprovado Marcelo Gordo (Universidade Federal do Amazonas) Aprovado Marina Anciães (Instituto Nacional de Pesquisas da Amazônia) Aprovado iii K11 CDD 19. ed. 597.8044 Kaefer, Igor Luis Diversificação evolutiva em anfíbios amazônicos (Allobates, Dendrobatoidea): caracteres genéticos, morfológicos e comportamentais / Igor Luis Kaefer.--- Manaus : [s.n.], 2012. xviii, 131 f. : il. color. Tese (doutorado) --- INPA, Manaus, 2012 Orientador : Albertina Pimentel Lima Coorientador : Izeni Pires Farias Área de concentração : Ecologia 1. Amazônia. 2. Anura. 3. Bioacústica. 4. Biogeografia. 5. DNA mitocondrial. 6. Filogeografia. I. Título. CDD 19. ed. 597.8044 Sinopse: Mecanismos históricos e geográficos responsáveis pela diversificação evolutiva foram avaliados em três espécies de Allobates ao longo de suas áreas de distribuição na Amazônia brasileira. Foram analisados aspectos como o grau de estereotipia e a plasticidade dos sinais sexuais. Foram estudadas as relações genealógicas, a história filogeográfica das populações, e a divergência relativa em caracteres genéticos (DNA mitocondrial), morfológicos (medidas morfométricas externas) e comportamentais (caracteres acústicos). Por fim, foi avaliado o papel de caracteres fenotípicos em atribuir o pertencimento de indivíduos a seus respectivos agrupamentos genéticos. Palavras-chave: Amazônia, Anura, bioacústica, biogeografia, DNA mitocondrial, filogeografia. iv AGRADECIMENTOS Sou grato à Albertina P. Lima não somente pela orientação, mas por ter me permitido e incentivado a procurar novos horizontes na pesquisa, muito além daqueles inicialmente traçados no projeto de tese. À Izeni P. Farias pelo incondicional apoio em todas as etapas e pelos ensinamentos em genética molecular que deverão fazer diferença ao longo do meu futuro acadêmico. Ao Pedro Ivo Simões e Luciana K. Erdtmann que, além de grandes amigos e colegas de laboratório, participaram em diversas etapas dessa caminhada e representam para mim exemplos a serem seguidos na conduta científica. Ao prof. Miguel Vences, que permitiu que eu concluísse a redação desta tese nas dependências da Technische Universität Braunschweig. Ao Instituto Nacional de Pesquisas da Amazônia (INPA). Primeiramente, por reunir pesquisadores que me inspiraram a conhecer mais sobre esse vasto bioma. Além disso, por fornecer infraestrutura e apoio logístico para a condução dos meus estudos. Ao Programa de Pós-Graduação (PPG) em Ecologia do INPA. Em especial, às coordenadoras, Flávia Costa e Claudia Keller, pela dedicação ao curso. Aos professores do Programa, cujo legado vai além do que nos foi apresentado durante as disciplinas. Aos colegas e amigos do PPG Ecologia, em especial aqueles ingressos em 2008. Hoje já nos encontramos “dispersos” pelo mundo. Porém, continuamos unidos pelo forte sentimento de amizade que se consolidou ao longo dos anos de convivência na floresta e na cidade. À Andresa Melo, Beverly Franklin, Érica Magalhães, Giselle Campos, Itamara da Gama, Rosirene Farias, Walmira da Paz, Maria Carmozina de Araújo e Rafael de Fraga por terem descomplicado diversos trâmites burocráticos e pelas conversas animadas na hora do café. O trabalho de campo, com amostragens pontuais ao longo da Amazônia Central, envolveu dezenas de famílias que gentilmente me acolheram em suas casas e cederam suas embarcações para a procura dos Allobates. Agradeço em especial à D. Irene Melo, Moisés S. Melo, Arthur O. Nascimento, Priscila P. Correa, Raimundo N. Amorim e aos diversos outros “Raimundos” que me auxiliaram nessa empreitada. Também agradeço aos motoristas Srs. Lourival dos Santos e Arnoldo Souza pelas idas e vindas aos sítios de coleta. Pela companhia v em campo, sou grato à Aída Rodrigues, Ana Carolina Mello, Anelise Montanarin, Bruno M. Tsuji-Nishikido, Francisco C. de Freitas e Maria Aparecida Carvalho. Pela linda e proveitosa viagem pelos Rios Negro e Unini, agradeço ao Niro Higuchi e toda a equipe do Laboratório de Manejo Florestal do INPA. Sou grato à Universidade Federal do Amazonas através do Laboratório de Evolução e Genética Animal (LEGAL) pela possibilidade de realizar as análises moleculares apresentadas nesta tese. À grande família do LEGAL, com a qual tanto pude aprender: Mário Nunes, Adriano Cantuária, Concy Santos, Deyla Oliveira, Fabíola Rodrigues, Maria Lizarazo, Gabriela de Pinho, Carla Bantel, Valéria Machado, Kelmer Passos, Jaqueline Fortuna, Roberta Canton, Jessica Souza, Roberta Oliveira, Natasha Meliciano, Olavo Colatreli, Rafaela Cardoso, Fabio Muniz, Adam Leão, Nicole Dutra, Marina Anciães e Tomas Hrbek. Em especial, à Waleska Gravena, Daniela Leroy e Edvaldo Mota por terem me acompanhado em meus passos iniciais dentro do lab. Aos pesquisadores que dedicaram seu tempo para avaliar uma ou mais das diversas etapas da execução deste projeto, desde sua proposição junto ao curso até a avaliação do documento final: Adrian Garda, Andrew Crawford, Camila Ribas, Heike Pröhl, Ilse Walker, Jeffrey Podos, Marcelo Gordo, Marcelo Menin, Marina Anciães, Mario Cohn-Haft, Maristerra Lemes, Vanessa Verdade e Walter Hödl. Agradeço também pelas construtivas contribuições recebidas de Adolfo Amézquita e Janet Reid, bem como aquelas provenientes dos editores e revisores anônimos dos periódicos aos quais os capítulos foram submetidos para publicação. Ao Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) pela bolsa de doutorado a mim concedida e pelo financiamento do projeto por meio dos processos CTAmazônia 553997/2006-8 e 575572/2008-6. Agradeço imensamente aos amigos e colegas de faculdade que, a despeito de todas as dificuldades decidiram, junto comigo, realizar o sonho de estudar na distante Amazônia: Anderson Bueno, Cristian Dambros, Daiani Kochhann e Mauro Ortiz. Estou certo de que nossa bravura rendeu e renderá excelentes frutos. À Sonia Cechin, da Universidade Federal de Santa Maria (UFSM), por ter me contagiado com sua paixão pelo estudo dos anfíbios e répteis. Aos colegas do Laboratório de Herpetologia da UFSM pelo incentivo e companheirismo que se estendem até os dias atuais. vi Aos amigos de longa data que, mesmo distantes fisicamente, têm me apoiado e continuado a fazer parte do meu dia-a-dia. Aos novos amigos que conquistei durante minha estadia em Manaus, e que acabaram por tornar-se minha família manauara. Muito obrigado por fazerem parte da minha vida! Espero tê-los comigo sempre! E, principalmente, agradeço à minha família por ter me apoiado em todas as minhas decisões, e por constantemente expressar seu orgulho pelos meus estudos. Felizmente, o amor que nos une independe de quaisquer barreiras físicas. vii “Ansioso, esperei o amanhecer: a natureza, aqui, além de misteriosa é quase sempre pontual." Milton Hatoum viii RESUMO Esse estudo objetivou elucidar mecanismos históricos e geográficos responsáveis pela diversificação evolutiva em três espécies de Allobates ao longo de suas áreas de distribuição na Amazônia brasileira. No primeiro capítulo, avaliou-se a potencial importância dos caracteres acústicos do canto de anúncio de Allobates paleovarzensis na discriminação entre indivíduos e populações. Além disso, foram acessados os efeitos de fatores individuais, ambientais e geográficos na variabilidade dos componentes deste sinal sexual. Nenhum dos caracteres acústicos analisados foi indicado como potencial sinal para reconhecimento social ou seleção sexual na espécie estudada. Lado do rio e distância geográfica não afetaram significativamente a variabilidade acústica, indicando que forças evolutivas estabilizadoras locais devem ser importantes no processo de diferenciação do sinal sexual. O segundo capítulo apresentou a caracterização do grau de variabilidade de cada componente acústico do sinal sexual entre populações e entre as espécies irmãs e alopátricas Allobates nidicola e A. masniger. Em adição, testaram-se os efeitos do tamanho corporal, da temperatura ambiental, e da distância geográfica na variabilidade acústica do sistema de estudo. Propriedades espectrais foram mais distintivas que propriedades temporais entre populações e entre espécies, com maior distinção ao nível interpopulacional quando comparado ao nível interespecífico. Houve acentuada plasticidade em caracteres temporais do canto de anúncio. O efeito do Rio Madeira como barreira foi significativo entre todas as variáveis acústicas analisadas. A ausência de efeitos de isolamento por distância indica que pressões seletivas estabilizadoras locais devem ser mais importantes que a deriva genética na evolução da diferenciação do canto de anúncio. No terceiro capítulo, investigaram-se mecanismos evolutivos e fatores geográficos levando à especiação, avaliando-se as relações genealógicas, a história filogeográfica e a divergência de caracteres genéticos, morfológicos e comportamentais entre populações das três espécies. Foi observada significativa estruturação genética populacional. A distribuição da variabilidade genética nos dois sistemas de estudo exibiu assinaturas congruentes com a transposição de grandes rios amazônicos. A efetividade do baixo Rio Madeira como uma barreira ao fluxo gênico foi maior do que a do médio Amazonas. Uma grande porção da diferenciação genética foi correlacionada com distância geográfica linear, evidenciando a importante atuação da deriva genética na diversificação do marcador mitocondrial estudado. Os caracteres acústicos e morfológicos não responderam às distâncias lineares entre populações. Em adição, análises de função discriminante falharam em classificar corretamente os indivíduos analisados em seus agrupamentos genéticos de acordo com seus fenótipos. Esse resultado é surpreendente levando-se em consideração a alta estruturação genética observada, e sugere que limitações evolutivas podem estar impedindo a diferenciação fenotípica. Em geral, a diversificação evolutiva refletiu efeitos de isolamento por distância e de barreiras vicariantes, e a contribuição relativa de cada fator diferiu entre os dois sistemas estudados. ix ABSTRACT Evolutionary diversification in Amazonian amphibians (Allobates, Dendrobatoidea): genetic, morphological and behavioral characters This study aimed to elucidate the historical and geographical mechanisms responsible for evolutionary diversification in three species of Allobates throughout their distributional areas in Brazilian Amazonia. In the first chapter, the potential importance of the acoustic characters of the advertisement call of Allobates paleovarzensis in discriminating among individuals and populations was evaluated. In addition, I assessed the effects of individual, environmental and spatial factors on the variability of the components of this sexual signal. None of the acoustic characters analized was indicated as a potential cue for social recognition or sexual selection in this species. River side and geographical distance did not significantly affect the acoustic variability, indicating that stabilizing local evolutionary forces might be important in the process of sexual signal differentiation. The second chapter presents the characterization of the degree of variability of each component of the acoustic sexual signal among populations and between the allopatric sister species Allobates nidicola and A. masniger. Additionally, I tested the effects of body size, environmental temperature, and geographical distance on the acoustic variability within the study system. Spectral traits were more effective than temporal properties in distinguishing among populations and between species, with greater difference at the interpopulation level when compared to the interspecific level. The temporal characters of the advertisement call were highly plastic. The effect of the Madeira River as a barrier was significant for all acoustic traits analyzed. The absence of isolation-by-distance effects indicates that local stabilizing selection pressures might be more important than genetic drift in the evolution of the advertisement call differentiation. In the third chapter, I investigated evolutionary mechanisms and geographical factors leading to speciation, by assessing the genealogical relationships, the phylogeographic history, and the divergence in genetic, morphological and behavioral traits among populations belonging to the three species. There was significant population genetic structure. The distribution of genetic variability in the two study systems exhibited signatures consistent with the transposition of large Amazonian rivers. The effectiveness of the lower Madeira River as a barrier to gene flow was higher than that of the middle Amazon. A large portion of the genetic differentiation was correlated with linear geographical distance, showing the important role of genetic drift in the diversification of the mitochondrial marker studied. The acoustic and morphological traits did not respond to linear distances between populations. In addition, discriminant function analyses failed to correctly classify the individuals analyzed in their genetic clusters based on their phenotypes. This result is surprising in view of the high genetic structure observed, and suggests that evolutionary constraints may be preventing phenotypic differentiation. In general, the evolutionary diversification reflected effects of isolation by distance and of vicariant barriers, and the relative contribution of each factor differed between the two systems studied. x SUMÁRIO LISTA DE TABELAS xi LISTA DE FIGURAS xvi INTRODUÇÃO GERAL 1 OBJETIVOS 4 Capítulo I – Artigo: Sexual signals of the Amazonian frog Allobates 5 paleovarzensis: geographic variation and stereotypy of acoustic traits Capítulo II – Artigo: Beyond the river: underlying determinants of population 30 acoustic signal variability in Amazonian direct-developing Allobates (Anura: Dendrobatoidea) Capítulo III – Artigo: The early stages of speciation in Amazonian forest 52 frogs: phenotypic conservatism despite strong genetic structure SÍNTESE 106 REFERÊNCIAS BIBLIOGRÁFICAS 109 ANEXOS 125 xi LISTA DE TABELAS Capítulo I Tabela 1. Efeito da temperatura e do tamanho corporal nas propriedades acústicas de 22 Allobates paleovarzensis (n = 123 machos). Coeficientes de determinação das regressões lineares (r2), estatísticas F e valores de significância (p) são apresentados. Para tamanho corporal (CRC), propriedades acústicas foram ajustadas com relação à temperatura. Valores de p significativos após correção de Bonferroni estão marcados com asteriscos. Tabela 2. Estatística descritiva de caracteres do canto de anúncio baseada em médias 23 gerais dos parâmetros acústicos calculadas a partir dos valores médios de 123 machos registrados em dez localidades ao longo da distribuição de Allobates paleovarzensis, representando variação natural entre 23.1 and 28.0 oC. Coeficientes de variação médio intraindividual, médio intrapopulacional, e geral (CVwi, CVwp, CVo, %) de dez propriedades do canto de anúncio (ajustadas para 25.4oC por meio de regressão linear) de Allobates paleovarzensis. As propriedades estão classificadas por tipo com base nos CVs médios intraindividuais de acordo com Gerhardt (1991). Tabela 3. Análise hierárquica de (co-)variância de caracteres do canto de anúncio 24 (variáveis dependentes) de Allobates paleovarzensis. População (aninhada em lado do rio) e lado do rio são as variáveis categóricas independentes. No modelo b, tamanho é usado como covariável. Valores de p significativos após correção de Bonferroni estão marcados com asteriscos. Capítulo II Tabela 1. Estatística descritiva dos caracteres acústicos analisados, com base em 43 médias individuais das propriedades do canto de 100 indivíduos machos, ajustadas para 25.9 °C utilizando-se regressão linear. Coeficientes de variação (%, apresentados como média ± DP) intrapopulacional (CVwp), intraespecífico (CVws), e geral (CVo) de seis propriedades do canto de anúncio de Allobates nidicola e A. masniger. Tabela 2. Efeito da temperatura e do tamanho corporal nos caracteres acústicos dos indivíduos estudados (n = 100 machos). Coeficientes de determinação ajustados das regressões lineares (r2), estatísticas F e valores de significância (p) são apresentados. Para tamanho corporal, estatísticas foram calculadas utilizando-se caracteres acústicos 44 xii ajustados para temperatura. Valores de p significativos após correção de Bonferroni estão marcados com asteriscos. Tabela 3. Análise hierárquica de (co-)variância de caracteres do canto de anúncio 45 (variáveis dependentes) de A. nidicola e A. masniger. População (aninhada em lado do rio) e lado do rio (espécies) são as variáveis categóricas independentes. No modelo b, tamanho é usado como covariável. Valores de p significativos após correção de Bonferroni estão marcados com asteriscos. Tabela 4. Testes de Mantel avaliando correlações entre distâncias acústica (DAco), de 46 tamanho corporal (DSVL), e geográfica (DGeo) de Allobates amostrados em dez localidades na Amazônia Central. Por causa dos efeitos do rio, os testes também foram restritos aos lados esquerdo e direito do Rio Madeira (cinco localidades cada). Valores de p significativos após correção de Bonferroni estão marcados com asteriscos. Capítulo III Tabela 1. Sítios de estudo na Amazônia brasileira com respectivos tamanhos 85 amostrais e coordenadas geográficas. Localidades tipo estão indicadas com asteriscos. Os tamanhos amostrais correspondem ao número de Allobates amostrados para cada classe de caracter. Tabela 2. Análises de variância molecular (AMOVA) baseadas em fragmentos do 86 DNAr 16S mitocondrial. A distribuição relativa da variabilidade genética é apresentada de acordo com os níveis hierárquicos de estruturação genética em cada sistema de estudo de Allobates. Tabela 3. Índices de fixação FST par-a-par (matriz inferior esquerda) e distâncias 87 genéticas médias Kimura-2-parâmetros (matriz superior direita). Os valores foram calculados entre as localidades de amostragem de cada sistema de estudo de Allobates na Amazônia brasileira. Valores de FST significativos estão marcados com asteriscos. Tabela 4. Estatística descritiva de parâmetros de polimorfismo genético e resultados de testes de neutralidade realizados em cada agrupamento genético de Allobates. Análise Bayesiana de Estruturação Genética foi utilizada para estimar a composição dos agrupamentos genéticos. n = tamanho amostral; h = número de haplótipos; S = número de sítios segregantes; π = distância média par-a-par (± um desvio padrão) entre amostras pertencentes ao mesmo agrupamento; D = D de Tajima; Fs = Fs de Fu; 88 xiii R2 = R2 de Ramos-Onsins & Rozas. Valores significativos em testes de neutralidade estão marcados com asteriscos. Tabela 5. Testes de expansão demográfica realizados de acordo com os agrupamentos 89 genéticos de Allobates indicados pela BAPS. Valores foram inferidos a partir da soma do desvio quadrado (SSD) entre distribuições de mismatch observadas e esperadas, e do índice de raggedness de Harpending (Hri). Os testes não foram conduzidos no agrupamento 4 de Allobates paleovarzensis devido ao pequeno número de amostras que o compõe. n = número de amostras. Os valores não significativos de p de ambos os testes indicam que a hipótese nula de expansão populacional através do tempo não pode ser rejeitada. Tabela 6. Testes de Mantel simples e parciais avaliando correlações entre distâncias 90 geográficas, fenotípicas e genéticas de Allobates entre as localidades estudadas na Amazônia brasileira. Testes de Mantel simples são apresentados como “Matriz 1” X “Matriz 2” e testes de Mantel parciais são apresentados como “Matriz 1” X “Matriz 2”.“Matriz Covariável”. GenD = distância genética; GeoD = distância geográfica; MorD = distância morfológica; AcoD = distância acústica; AcoDwa = Distância acústica sem ajuste de tamanho corporal; River = lado do rio (variável binária). Correlações significativas são indicadas com asteriscos. Recurso Online 1. Medidas do canto de anúncio de indivíduos de Allobates paleovarzensis em cada localidade de estudo na Amazônia brasileira. Valores são apresentados como média (acima) e desvio padrão (abaixo). Os caracteres do canto são: Repetição da nota (NR, em notas/s); Duração da nota (ND, em s); Intervalo entre notas, como o intervalo silencioso entre duas notas consecutivas de um canto (InI, em s); Repetição do canto (CR, em cantos/s); Duração do canto (CD, em s); Intervalo entre cantos, como o intervalo silencioso entre dois cantos consecutivos (IcI, em s); Frequência máxima (pico), como a frequência de maior intensidade calculada para toda a nota por meio da função power spectrum do programa Raven Pro 1.3 (MF, em Hz); Frequência mais baixa (LF, em Hz); Frequência mais alta (HF, em Hz); Modulação da nota, como a diferença entre as frequências mais alta e mais baixa da nota (NM, em Hz). Os códigos das localidades de amostragem, os tamanhos amostrais e as respectivas coordenadas geográficas são apresentados na Tabela 1. 98 xiv Recurso Online 2. Medidas do canto de anúncio de indivíduos de Allobates nidicola 99 e A. masniger em cada localidade de estudo na Amazônia brasileira. Valores são apresentados como média (acima) e desvio padrão (abaixo). Os caracteres do canto são: Duração da nota (ND, em s); Intervalo entre notas (IN, em s); Frequência mais baixa (LF, em Hz); Frequência mais alta (HF, em Hz); Modulação da nota, calculada como a diferença entre as frequências mais alta e mais baixa (NM, em Hz); e Frequência pico (PF, em Hz). Os códigos das localidades de amostragem, os tamanhos amostrais e as respectivas coordenadas geográficas são apresentados na Tabela 1. Recurso Online 3. Medidas morfométricas (em mm) de indivíduos de Allobates 100 paleovarzensis em cada localidade de estudo na Amazônia brasileira. Valores são apresentados como média (acima) e desvio padrão (abaixo). Os caracteres morfométricos são: Comprimento rostro-cloacal (SVL); Comprimento da cabeça, da articulação da maxila até a ponta do focinho (HL); Largura da cabeça ao nível da articulação da maxilla (HW); Comprimento do focinho, do canto anterior do olho até a ponta do focinho (SL); Distância do olho à narina, do canto anterior do olho ao centro da narina (EN); Distância entre narinas (IN); Comprimento do olho, do canto anterior ao posterior (EL); Distância interorbital (IO); Diâmetro do tímpano (TYM); Comprimento do antebraço, da extremidade proximal do tubérculo palmar até a borda exterior do cotovelo flexionado (FAL); Comprimento do braço, da sua inserção no tronco até a borda exterior do cotovelo flexionado (UAL); Comprimento da mão, da extremidade proximal do tubérculo palmar até a ponta dos dedos I, II, III e IV (HAND1, HAND2, HAND3 e HAND4); Largura do disco no dedo III da mão (WFD); Comprimento da tíbia, da borda externa do joelho flexionado até o calcanhar (TL); Comprimento do pé, da extremidade proximal do tubérculo metatarsal até a ponta do dedo IV (FL); Comprimento do fêmur (LL); Diâmetro do tubérculo palmar (DPT); Largura do tubérculo tenar (WTT); Largura do disco no dedo IV do pé (WTD); Largura do dedo III da mão (WPF). Os códigos das localidades de amostragem, os tamanhos amostrais e as respectivas coordenadas geográficas são apresentados na Tabela 1. Recurso Online 4. Medidas morfométricas (em mm) de indivíduos de Allobates nidicola e A. masniger em cada localidade de estudo na Amazônia brasileira. Valores são apresentados como média (acima) e desvio padrão (abaixo). As medidas morfométricas são descritas no Recurso Online 3. Os códigos das localidades de amostragem, os tamanhos amostrais e as respectivas coordenadas geográficas são apresentados na Tabela 1. 101 xv Recurso Online 5. Distribuição dos haplótipos de rDNA 16S de Allobates entre 20 102 localidades amostradas na Amazônia brasileira. Números de coleta de espécimes testemunho (INPA-H) e números de acesso do GenBank são apresentados. As localidades amostradas estão numeradas de acordo com a Tabela 1. Recurso Online 7. Matriz de classificação da análise de função discriminante, 104 baseada em dados fenotípicos de machos de Allobates paleovazensis pertencentes a quarto grupos genéticos. Grupos estão codificados de C1 a C4, de acordo com a Tabela 4. Caracteres de forma corporal foram usados para discriminar indivíduos morfologicamente. Medidas acústicas foram ajustadas para temperatura e para temperatura e tamanho corporal simultaneamente por meio de regressões lineares. O número e a percentagem de indivíduos corretamente atribuídos a cada grupo genético são indicados. Recurso Online 8. Matriz de classificação da análise de função discriminante, baseada em dados fenotípicos de machos de Allobates nidicola e A. masniger pertencentes a sete grupos genéticos. Grupos estão codificados de C1 a C7, de acordo com a Tabela 4. Caracteres de forma corporal foram usados para discriminar indivíduos morfologicamente. Medidas acústicas foram ajustadas para temperatura e simultaneamente para temperatura e tamanho corporal por meio de regressões lineares. O número e a percentagem de indivíduos corretamente atribuídos a cada grupo genético são indicados. 105 xvi LISTA DE FIGURAS Capítulo I Figura 1. Distribuição geográfica das localidades de amostragem na Amazônia 26 Central. 1 = Careiro, 2 = Janauacá, 3 = Manaquiri, 4 = Hiléia, 5 = Novo Airão, 6 = Anamã, 7 = Anori, 8 = Codajás, 9 = Unini, 10 = Barcelos. Figura 2. (A) Oscilograma e (B) espectrograma de três cantos de anúncio multinotas, 27 e (C) espectro de força de uma única nota de um macho de Allobates paleovarzensis em sua localidade tipo em Careiro da Várzea, Brasil. Figura 3. Coeficientes de variação intraindividuais (CVwi) para dez propriedades 28 acústicas do canto de anúncio de Allobates paleovarzensis. Retângulos cinza indicam os 50% centrais dos dados (interquartil), e linhas horizontais representam os valores medianos. Pontos representam valores discrepantes. Linhas verticais indicam o 10º e o 90º percentis. Figura 4. Variação em propriedades do canto (ajustadas para 25.4 oC por meio de 29 regressão linear) e tamanho do corpo ao longo da distribuição de Allobates paleovarzensis na Amazônia Central. Retângulos cinza indicam os 50% centrais dos dados (interquartil), e linhas horizontais representam os valores medianos. Pontos representam valores discrepantes. Linhas verticais indicam o 10º e o 90º percentis. Capítulo II Figura 1. Machos de A. nidicola (acima) e A. masniger (abaixo) vocalizando em suas 48 respectivas localidades tipo na Amazônia brasileira. Figura 2. Localidades de amostragem visitadas neste estudo: 1 Km 12 da Estrada de 49 Autazes, localidade tipo de A. nidicola; 2 BR-319, km 260; 3 PPBio Manaquiri; 4 BR319 Tupana; 5 Vila Gomes; 6 Borba; 7 Novo Aripuanã; 8 Estrada para Apuí; 9 Jacareacanga; 10 PARNA da Amazônia, localidade tipo de A. masniger. Figura 3. Representação gráfica das medidas acústicas adotadas neste estudo. Oscilograma (a) e espectrograma (b) de uma sequência de cantos de um macho topotípico de A. masniger (CRC = 18.5 mm; INPA-H 28196), registrado em temperatura do ar de 23.7 °C. Espectro de força (c) de um único canto deste indivíduo. As frequências mais alta e mais baixa foram medidas a 20 dB abaixo da intensidade 50 xvii pico (linha pontilhada horizontal), valor no qual a energia do sinal pode ser claramente distinguida do ruído de fundo. A modulação da nota foi calculada como a diferença entre as frequências mais alta e mais baixa do canto. Figura 4. Coeficientes de variação intrapopulacionais (CV, em percentagem) para 51 seis propriedades acústicas do canto de anúncio dos indivíduos estudados. Retângulos cinza indicam os 50% centrais dos dados (interquartil), e linhas horizontais representam os valores medianos. Linhas verticais indicam o 10º e o 90º percentis. Note que os coeficientes de variação mais baixos são relacionados a caracteres espectrais do canto. Capítulo III Figura 1. Localização geográfica dos pontos de amostragem (A) e rede de haplótipos 92 (B) para Allobates paleovarzensis. A rede foi construída a partir de 142 sequências de DNAr 16S. O tamanho e a cor de cada elipse indicam a frequência e origem geográfica dos indivíduos que apresentam aquele haplótipo. Pontos brancos e barras transversais representam haplótipos intermediários não amostrados (faltantes). As localidades amostradas estão numeradas de acordo com a Tabela 1. Figura 2. Localização geográfica dos pontos de amostragem (A) e rede de haplótipos 93 (B) para Allobates nidicola e A. masniger. A rede foi construída a partir de 134 sequências de DNAr 16S. O tamanho e a cor de cada elipse indicam a frequência e origem geográfica dos indivíduos que apresentam aquele haplótipo. Pontos brancos e barras transversais representam haplótipos intermediários não amostrados (faltantes). As localidades amostradas estão numeradas de acordo com a Tabela 1. Figura 3. Gráfico de barras da Análise Bayesiana de Diferenciação Genética. (A) 142 sequências individuais de DNAr 16S de Allobates paleovarzensis de dez localidaes de amostragem. (B) 134 sequências individuais de DNAr 16S de Allobates nidicola e A. masniger de outras dez localidades. Cores distintas representam cada grupo genético estimado. Indivíduos são exibidos de acordo com as localidades de amostragem, as quais estão numeradas de acordo com a Tabela 1. 94 xviii Figura 4. Árvore datada dos sistemas (A) paleovarzensis e (B) nidicola-masniger 95 baseada em fragmentos do gene DNAr 16S. Somente haplótipos únicos foram incluídos. Estimativas de idade e intervalos de confiança (barras horizontais) das divergências foram obtidos via BEAST (Bayesian Evolutionary Analysis Sampling Trees). As probabilidades posteriores dos nós mais antigos são apresentadas. Terminais foram coloridos de acordo com suas localidades de origem, as quais estão representadas nas Figuras 1 e 2. Terminais pretos e brancos são grupos externos. As barras verticais pretas e cinzas indicam a origem de cada haplótipo de acordo com o lado do rio. Recurso Online 6. Distribuições de mismatch obtidas a partir de diferenças par-a-par de sítios nucleotídicos entre sequências de DNAr 16S que compõem os grupos genéticos dos sistemas (A) paleovarzensis e (B) nidicola-masniger. Os grupos foram determinados via Análise Bayesiana de Diferenciação Genética. Diferenças par-a-par não foram calculadas para o Grupo 4 de Allobates paleovarzensis devido ao pequeno número de amostras que o compõe. 103 1 INTRODUÇÃO GERAL A integração entre distintas áreas do conhecimento tem permitido um melhor entendimento a respeito de padrões e mecanismos de diferenciação entre populações de diversos organismos (Crisci et al., 2003; Santini et al., 2012). A análise conjunta de diferentes classes de caracteres permite elucidar mecanismos evolutivos levando à especiação, uma vez que a história inferida a partir de sequências de DNA proporciona um contexto demográfico histórico para avaliar a divergência em diversos atributos fenotípicos (Wiens, 2008; Campbell et al., 2010; Guillot et al., 2012). Importantes avanços podem ser obtidos por meio da investigação do padrão geográfico no qual a variabilidade de uma determinada classe de caracteres é distribuída entre populações, assim como da ordem em que novidades evolutivas aparecem (West-Eberhard, 1989; Dall, 1997; Panhuis et al., 2001). Tal cenário multicaráter tem impulsionado esforços para a investigação das forças envolvidas na promoção e manutenção da diversificação intraespecífica e do processo de formação de novas espécies. Em adição à variedade de classes de marcadores, estudos considerando diferentes planos organizacionais contribuem significativamente para a compreensão da diversificação evolutiva, uma vez que esta pode ser acessada a partir de distintos níveis da hierarquia biológica (Castellano et al., 2002; Diniz-Filho et al., 2008). Assim, espera-se que estudos dirigidos à variabilidade de caracteres aos níveis de indivíduo, população e espécie possam proporcionar respostas complementares a respeito da multiplicidade de forças evolutivas envolvidas nesse processo. A compreensão dos processos de especiação requer o estudo de fatores geográficos que podem sustentar a origem e diversificação de linhagens, tais como vicariância e isolamento por distância (Hutchison & Templeton, 1999; Zink et al., 2000). Uma vez que efeitos de isolamento por distância são inferidos a partir de correlações entre distâncias geográficas e de caracteres fenotípicos ou genotípicos, estas podem ser afetadas pela presença de rios como barreiras, os quais podem levar a dissimilaridades desproporcionalmente maiores do que aquelas produzidas pela distância linear. Na Amazônia, investigações sobre o padrão de distribuição de espécies (e.g., Wallace, 1852; Haffer, 1969; Ayres & CluttonBrock, 1992; Hayes & Sewlal, 2004) e avaliações genealógicas em fina escala (e.g., CohnHaft, 2000; Funk et al., 2007; Solomon et al., 2008; Amézquita et al., 2009) têm inferido o efeito desses dois fatores históricos em organismos terrestres, embora a contribuição relativa atribuída a cada fator tenha variado. Assim, até o presente momento, nenhum padrão 2 geográfico ou força paleoambiental emergiu como uma explicação geral para a diversificação da maioria dos grupos animais (Antonelli et al., 2010; Ribas et al., 2011). Nesse contexto, anfíbios anuros figuram como excelentes modelos de estudo em variação geográfica. A alta abundância dos anfíbios propicia amostragens densas ao longo de suas áreas de distribuição, e o sinal sexual dos anuros (canto de anúncio) permite a utilização de caracteres acústicos como marcadores populacionais (Wycherley et al., 2002; Lougheed et al., 2006). No nível individual, caracteres acústicos são utilizados pelos anuros para reconhecimento e escolha sexual (Gerhardt & Huber, 2002). Diferenças nesses sinais podem aumentar dramaticamente o potencial evolutivo para especiação devido à possibilidade de isolamento pré-zigótico (Panhuis et al., 2001; Pröhl et al., 2006; Boul et al., 2007). A natureza multidimensional dos sinais sexuais torna os caracteres acústicos potencialmente sujeitos a diferentes determinantes (Gerhardt, 1991; Prestwich, 1994; Gerhardt & Huber, 2002). Assim, ao decompormos esses sinais em suas propriedades, podemos medir o grau de variabilidade de caracteres quantitativos, e acessar seu potencial para seleção sexual e reconhecimento social, assim como entender o papel de determinados caracteres acústicos em processos microevolutivos (e.g., Friedl & Klump, 2002; Gasser et al., 2009; Bee et al., 2010). Entretanto, estudos bioacústicos comparativos devem considerar covariáveis que afetam o sinal sexual, mas que não necessariamente carregam sinal filogenético. Visto que caracteres acústicos apresentam plasticidade induzida pela temperatura ambiental e pelo tamanho do corpo (Prestwich, 1994), faz-se necessária a detecção e a consideração desses efeitos para a obtenção de medidas mais precisas de diferenciação acústica quando comparamos indivíduos, populações ou espécies (Castellano et al., 2002). Pelo fato de muitas vezes serem prontamente perceptíveis, caracteres morfológicos têm servido historicamente como os marcadores mais amplamente utilizados para o estudo de variação entre populações. Tal característica conferiu grande valor à morfologia no desenvolvimento de conceitos e diagnoses em nível específico (Cronquist, 1978; Coyne & Orr, 2004). Atualmente, a importância relativa desses caracteres em estudos evolutivos tem diminuído em face do desenvolvimento de outros marcadores, em especial os de origem molecular (Avise, 2004). Entretanto, uma abordagem interdisciplinar não deve prescindir da utilização de dados morfológicos, quando disponíveis, para a identificação de processos evolutivos e, frequentemente, para a proposição de rearranjos taxonômicos decorrentes de análises em nível populacional (Dayrat, 2005; Buckley, 2009). 3 Ao longo das últimas décadas, testemunhou-se um desenvolvimento expressivo no estudo do arranjo espacial de linhagens genéticas, com ênfase no entendimento de como o isolamento genético no tempo e espaço desencadeia eventos de especiação e cria assinaturas em larga escala na diversidade genética dentro de espécies e em conjuntos de espécies proximamente relacionadas (Avise, 2009). Desde o cunho do termo filogeografia em 1987 (Avise et al., 1987), essa disciplina tem sido beneficiada por grandes avanços metodológicos e conceituais na investigação de processos ecológicos e históricos que influenciam e explicam a distribuição geográfica de organismos (Buckley, 2009; Avise, 2009). Atualmente, esta área do conhecimento vem desenvolvendo métodos sofisticados dentro de um sistema rigoroso em termos de obtenção de marcadores e de ferramentas estatísticas para o teste de hipóteses definidas a priori (e.g., Nielsen & Beaumont, 2009; Emerson et al., 2010). Assim, a filogeografia tem promovido uma verdadeira revolução no estudo da biodiversidade por tornar-se um componente fundamental em biogeografia, e por estabelecer uma conexão entre a investigação de processos micro e macroevolutivos, atuando assim como uma extensão geográfica e temporal da genética de populações (Riddle & Hafner, 2004; Riddle, 2009). Nos três capítulos apresentados a seguir, procuro elucidar mecanismos evolutivos e fatores geográficos levando à diversificação evolutiva em dois sistemas de estudo. Para isso, associei a utilização de diferentes classes de marcadores em um contexto que contemplasse os níveis organizacionais intraindividual, intraespecífico e interespecífico. Como modelos de estudo, utilizei três espécies recentemente descritas de Allobates, anuros diurnos e terrestres pertencentes à família Aromobatidae (Grant et al., 2006): Allobates paleovarzensis Lima et al. (2010), Allobates nidicola (Caldwell & Lima, 2003) e Allobates masniger (Morales, 2002). Os estudos aqui apresentados constituem as primeiras avaliações de cunho comportamental e/ou evolutivo envolvendo estas espécies. Foram amostradas 20 populações de Allobates em duas áreas adjacentes da bacia amazônica, nos estados brasileiros do Amazonas e Pará, em uma região caracterizada por grandes porções de florestas tropicais. Estas áreas compreendem seções de dois grandes rios do bioma, o Amazonas e o Madeira, possibilitando assim avaliar a contribuição relativa desses rios e do isolamento por distância como barreiras biogeográficas para os organismos estudados. As hipóteses apresentadas em cada capítulo são testadas por meio de enfoques populacionais utilizando marcadores genéticos, morfológicos e comportamentais. Assim, o conjunto dos capítulos objetiva uma abordagem integrativa e transdisciplinar para as questões apresentadas. 4 OBJETIVOS Os objetivos gerais de cada capítulo foram os seguintes: Capítulo I – Acessar a potencial importância dos caracteres acústicos do canto de anúncio de Allobates paleovarzensis na discriminação entre indivíduos e populações ao longo da área de ocorrência da espécie, e avaliar o efeito de fatores individuais, ambientais e geográficos na variabilidade dos componentes do sinal sexual. Capítulo II – Caracterizar o grau de variabilidade de cada componente acústico do sinal sexual entre populações e entre espécies irmãs e alopátricas (Allobates nidicola e A. masniger), e testar o efeito do tamanho corporal, da temperatura ambiental, e da distância geográfica na variabilidade acústica do sistema de estudo. Capítulo III – Elucidar mecanismos evolutivos e fatores geográficos que levam à especiação, avaliando as relações genealógicas, a história filogeográfica e a divergência de caracteres genéticos, morfológicos e comportamentais entre populações de três espécies de anfíbios anuros da Amazônia Central. 18 Capítulo 1 Kaefer, I.L. & Lima, A.P. 2012. Sexual signals of the Amazonian frog Allobates paleovarzensis: geographic variation and stereotypy of acoustic traits. Behaviour 149:15-33. 6 1 2 3 Sexual signals of the Amazonian frog Allobates paleovarzensis: geographic variation and stereotypy of acoustic traits 4 5 6 Short title: Acoustic signal variability in Amazonian frogs Igor Luis Kaefer1) & Albertina Pimentel Lima 7 8 Coordenação de Pesquisas em Ecologia, Instituto Nacional de Pesquisas da Amazônia, CP 9 478, 69011–970, Manaus, Amazonas, Brazil. 10 1) Corresponding author’s e-mail address: [email protected] 11 12 Summary 13 Because of its close relationship with the process of evolutionary differentiation, it is expected 14 that geographic variability in acoustic sexual traits should be greater among than within 15 populations. This is particularly expected in organisms with typically high population genetic 16 structure and low dispersal abilities, such as anuran amphibians. We studied the acoustic traits 17 of the advertisement call in the small-sized dendrobatoid frog Allobates paleovarzensis 18 through its range in Central Amazonia. We accessed the variability of call traits from the 19 within-male to the among-population levels, and evaluated the degree of stereotypy of the call 20 characteristics. Call variability had comparable magnitudes within and among populations, 21 and was independent of the degree of stereotypy of call measurements. Therefore, none of the 22 call traits stood out as a potential cue for discrimination between populations. Spectral call 23 measurements were static and strongly related with body size, which explained between 30 24 and 35% of the variation of these acoustic traits. Temporal characters of the notes were 25 dynamic and influenced by environmental temperature (e.g., 27% of note rate variation), 26 whilst temporal measurements of the entire calls were not related to the co-factors analysed. 27 Both spectral and temporal call traits varied among populations and between sides of the 28 Amazon River. Our results also indicate that body size and sampling site jointly affected the 29 variability of the call traits. However, geographic distances among populations and the river 7 30 barrier had no significant effect on the overall acoustic variation, indicating that local 31 stabilising selective forces may be important in the process of call differentiation. 32 Key words: acoustic communication, advertisement call, Amazon River, biogeography, 33 Dendrobatoidea, intraspecific variation, signal stereotypy. 34 35 Introduction 36 Variability in secondary sexual characters is a pivotal component that underlies the 37 evolutionary framework in which selective forces take place, and changes in mate recognition 38 systems may even precede the evolution of other indicators of specific status (Ryan et al., 39 1996; Panhuis et al., 2001). Therefore, analyses of geographic variation in intraspecific 40 communication systems have the potential to detect the development of differentiation among 41 populations, allowing researchers to derive testable predictions about the evolutionary 42 significance of such divergences (Gerhardt & Huber, 2002). 43 An acoustic signal conveys information about its sender, such as species identity, body 44 condition, and mate quality, and these characteristics have been decoded from the signals in a 45 variety of animal groups, such as monkeys (Fischer et al., 2004), bats (Behr et al., 2007), 46 crickets (Verburgt et al., 2011) and frogs (Chakraborty et al., 2010). Given that these signals 47 are often related to sexual recognition and other aspects of pre-mating reproductive isolation 48 (Gerhardt & Huber, 2002), their decomposition into their spectral and temporal properties 49 should provide clues regarding their potential in sexual discrimination among populations 50 (e.g., Pröhl et al., 2007; Gomez et al., 2011) and even individuals (e.g., Gasser et al., 2009; 51 Bee et al., 2010; Melendez & Feng, 2010). Bioacoustic traits of individuals can be classified 52 as static or dynamic according to the continuum of coefficients of variation observed 53 (Gerhardt, 1991). Previous results suggested that female frogs impose more stabilising 54 selection on stereotyped (static) properties, while directional selection should be imposed on 55 more dynamic call traits, which show higher coefficients of variation (Gerhardt, 1991). In 56 addition, the degree of stereotypy has been proposed to be related to the spatial pattern of 57 variation of the different properties of the call (Castellano et al., 2000). 58 Spatial separation is a potential predictor of both gene flow and degree of 59 environmental similarity between populations (Lougheed et al., 2006). Geographic signal 8 60 differentiation is frequently observed in frogs, on both small (e.g., Boul et al., 2007; 61 Rodríguez et al., 2010) and large scales (e.g., Heyer & Reid, 2003; Amézquita et al., 2009). 62 Empirical patterns of spatial distribution of acoustic intraspecific variability have been 63 attributed to different evolutionary forces which could act synergistically: correlations 64 between acoustic and geographic distances are associated with neutral evolutionary processes 65 such as genetic drift (Pröhl et al., 2007; Amézquita et al., 2009), while vicariant effects are 66 invoked when differentiation in calls is congruent with the crossing of geographic barriers 67 (Simões et al., 2008; Rodríguez et al., 2010). Intraspecific signal divergence may reflect 68 adaptation to structural features of the environment that differ between habitats (Ryan et al., 69 1990; Ryan & Wilczynski, 1991), or may be associated with female preferences; the latter 70 may be interpreted as evidence of speciation caused by sexual selection (Boul et al., 2007; 71 Guerra & Ron, 2008). 72 As a highly energy-demanding activity, the emission of advertisement calls has 73 morphological and environmental constraints acting on ecological time scales, which are 74 regarded as co-factors and should be controlled in comparative bioacoustic studies (Prestwich, 75 1994). They constitute an additional source of variation, of which the extent of interference 76 varies between species and among the different acoustic traits considered (Gerhardt, 1991; 77 Gerhardt & Huber, 2002). Studies observed a correspondence between the degree of call trait 78 variability and the co-factors responsible for their variation (e.g., Castellano & Giacoma, 79 1998; Castellano et al., 2000), and our knowledge on the evolution of call trait stereotypy 80 should be benefited by a better comprehension of this association. 81 Because of its close relationship – either as cause or outcome – with the process of 82 evolutionary differentiation, it is expected that geographic variability in acoustic sexual traits 83 should be greater among than within populations (e.g., Castellano et al., 2002; Smith & 84 Hunter, 2005). This is particularly expected in organisms with typically high population 85 genetic structure and low dispersal abilities, such as many anuran amphibians (Lougheed et 86 al., 2006; Vences & Wake, 2007). The potential for a given call trait to act as a cue for 87 discrimination (through either social recognition or sexual selection) can be evaluated by 88 comparing the ratios between among- and within-population coefficients of variation (Pröhl et 89 al., 2007). Therefore, we analysed the variability in advertisement call traits throughout the 90 distribution of the Amazonian lowland frog Allobates paleovarzensis (Dendrobatoidea: 91 Aromobatidae) in order to access the variability of sexual signal in different levels. These are 9 92 small-sized, diurnal, leaf-litter frogs that inhabit seasonally flooded forests of Central 93 Amazonia (Lima et al., 2010). Dendrobatoid frogs are frequently used as models in both 94 exploratory and experimental studies of acoustic communication, although no such studies 95 have been conducted with any member of the dull-colored A. trilineatus species group (sensu 96 Morales, 2002). Related with comparing the within- and among-population variability in 97 acoustic sexual traits, we aimed to evaluate the degree of stereotypy of each call trait by 98 assessing coefficients of variation for individual frogs, test the association between 99 environmental temperature and male body size with acoustic characters, and test for the effect 100 of a river barrier leading to call differentiation in the study system. 101 102 Material and methods 103 Data acquisition 104 We recorded advertisement calls of Allobates paleovarzensis in ten localities (Figure 105 1), which will be regarded as individual populations in this study. Sampling sites were located 106 in the state of Amazonas, Brazil, on both sides of the Amazon River, and were at least 28 km 107 distant from each other. This study extends the known distribution of this species, for which 108 the confirmed occurrence was formerly restricted to its type locality in the municipality of 109 Careiro da Várzea (Lima et al., 2010). Therefore, this study comprises the entire known range 110 of A. paleovarzensis. Fieldwork was conducted between January 2009 and June 2010. In total, 111 we recorded advertisement calls of 123 males, with a mean of 12 males per population (SD = 112 3.8, range = 2–15; 12–15 when excluding Unini). Each male had its advertisement call 113 recorded for about three minutes with a Marantz PMD660 digital recorder (44.1 kHz 114 sampling rate; 16-bit resolution) and a Sennheiser K6/ME66 directional microphone 115 positioned approx. 1 m from the calling site. After each recording, we measured 116 environmental (air) temperature with a digital thermometer, and captured the frog, to measure 117 body size (snout-vent length, SVL, in mm) with a digital caliper. Voucher specimens were 118 deposited in the Herpetological Collection of the Instituto Nacional de Pesquisas da 119 Amazônia, in Manaus, Brazil. 120 Call measurements were obtained from recordings in Raven Pro 1.3 software (Charif 121 et al., 2008). The advertisement call of Allobates paleovarzensis consists of groups of single 10 122 notes separated by silent intervals (Lima et al., 2010; see Figure 2). For each recording, we 123 chose 10 calls with less background noise for detailed analyses. Note measurements, 124 including the spectral traits, were taken from the middle or next note of each call. Spectral 125 measurements were taken after a Fast Fourier Transform with frequency resolution of 82 Hz 126 and 2048 points. We considered ten call traits in our analyses: Note rate (NR, in notes/s); 127 Note duration (ND, in s); Internote interval, as the silent interval between two consecutive 128 notes of a call (InI, in s); Call rate (CR, in calls/s); Call duration (CD, in s); Intercall interval, 129 as the silent interval between two consecutive calls (IcI, in s); Maximum (peak) frequency, as 130 the frequency of higher intensity calculated for an entire note by a power spectrum function of 131 Raven Pro 1.3 (MF, in Hz); Lowest frequency (LF, in Hz); Highest frequency (HF, in Hz); 132 Note modulation, as the difference between the highest and the lowest frequencies of the call 133 (NM, in Hz). LF and HF were measured at 20 dB below the peak intensity, the value at which 134 the signal energy could still be clearly distinguished from background noise (Lima et al., 135 2010). 136 137 Data analysis 138 The data set was tested for normality. When appropriate, we log-transformed data 139 prior to carrying out the parametric statistical analyses. Given that certain call traits are 140 affected by environmental temperature and male body size (Gerhardt & Huber, 2002), we 141 tested to what extent these two co-variables affect each of the call parameters in the species 142 studied. For temperature, linear regression analyses were conducted. For all within- and 143 among-population analyses we used temperature-adjusted acoustic data, by calculating overall 144 regression coefficients (b) for all call parameters at the mean temperature of 25.4 oC (SD = 145 1.08, range = 23.1–28.0). We applied the equation Y adj = y – (b*tcalling site) + (b*tmean) where 146 Y adj = adjusted value of call measurement, b = regression coefficient and t = temperature. To 147 test for morphological correlates of call traits, we performed linear regression analyses 148 between body size and temperature-adjusted call measurements. The adjusted r2 values 149 obtained were used to estimate the percentage of variation explained by the independent 150 variable. 151 We calculated coefficients of variation (CV = 100%*SD/ ) for each call trait within 152 individuals, within populations, and over all individuals combined. For each focal male, we 153 determined the individual mean ( wi) and standard deviation (SDwi) of each call property (n = 11 154 10 calls/male), and we used these values to calculate a within-individual (i.e., within155 recording) coefficient of variation (CVwi = 100%*SDwi/ wi) (Bee et al., 2010). By averaging 156 all within-individual CVs, we obtained the mean within-individual CV (CVwi). This measure 157 allows us to assess the degree of stereotypy of each call property: call traits with a mean CVwi 158 less than 5% were considered static; those with a mean CVwi greater than 12% were 159 considered dynamic; and those with a mean CVwi between 5% and 12% were classified as 160 intermediate (Gerhardt, 1991). In addition, we calculated within-population CVs based on the 161 mean ( wp) and standard deviation (SDwp) over all call-recorded individuals in each 162 population (CVwp = 100%*SDwp/ wp). By averaging all within-population CVs, we obtained 163 the mean within-population CV (CVwp) (Pröhl et al., 2007). Overall CVs (CVo) were obtained 164 from the grand mean and standard deviation based on averaging all the individual means from 165 our sample (CVo = 100%*grand SD/grand ). This calculation has also been regarded as 166 among-individual (Bee et al., 2010) and among-population coefficients of variation (Pröhl et 167 al., 2007). The Unini locality was excluded from the analysis of geographic variation because 168 of the small sample size. Based on these measurements, we determined the ratio between the 169 overall and mean within-population variation as CVo/CVwp. 170 To compare the difference in calls (with temperature-adjusted traits) between sides of 171 the Amazon River and among populations, nested analyses of variance (ANOVA, populations 172 nested on river sides) were conducted. We used two different models for each call trait: (a) a 173 nested ANOVA without co-factor and (b) a nested ANOVA with size as co-factor, thus 174 constituting a nested ANCOVA. These two approaches were used in order to compare the F 175 statistics obtained through the inclusion of body size as a co-factor in the analyses: if an 176 interaction occurs between co-factors and populations (sampling sites), we would expect a 177 decline in F statistics due to worsening of the model. All statistical analyses, unless otherwise 178 stated, were conducted in SYSTAT 12.0 (Wilkinson, 2007). 179 We tested the correlation between acoustic/body size distances among sampling sites 180 and respective linear geographic distances by applying Mantel tests. In addition, partial 181 Mantel tests were conducted to check for correlations between acoustic/body size distances 182 and their division by the Amazon River channel, while controlling for effects of linear 183 geographic distance among populations. For this, we constructed a binary correspondence 184 matrix, designating the values “0” and “1” for localities within the same and between opposite 185 sides of the Amazon River, respectively. Acoustic distance matrices were obtained from the 186 dataset of temperature-adjusted call measurements by calculating pairwise Euclidean 12 187 distances between their scores on first and second components (which together explained 188 79.0% of the total acoustic variation) produced by a principal components analysis (PCA). As 189 the scores of the first principal component were correlated with body size (linear regression r2 190 = 0.51, F = 9.39, p = 0.017), we regressed them against corresponding mean SVL 191 measurements for each population, and used the residuals as new, size-independent acoustic 192 variables from which new Euclidean distances were calculated. Mantel tests were conducted 193 in ZT (Bonnet & van de Peer, 2002) using permutation of the null models (Anderson & 194 Legendre, 1999), and applying 10000 randomisations. 195 196 Results 197 Effect of temperature and body size on acoustic traits 198 There was a high temperature dependency of all temporal traits related to the notes of 199 the advertisement call. This relation was not observed in temporal call traits related to the 200 entire call, nor in spectral characters (Table 1). Males produced shorter and more closely 201 spaced notes with increasing temperature. Environmental temperature affected one spectral 202 property (note modulation), which was highly correlated with the temporal trait of note 203 duration (rs = 0.53, p < 0.001, n = 123). The size of the frogs was related only to spectral 204 parameters of the advertisement call. There were significant negative relationships between 205 body size and the maximum, lowest and highest frequencies. Male SVL explained between 30 206 and 35% of the variation in spectral traits (Table 1). 207 208 Stereotypy of acoustic traits 209 According to the within-male coefficients of variation observed, the most dynamic 210 traits were related to temporal characteristics of the call. Call duration (range = 0.75–8.73 s) 211 and inter-call interval (range = 1.02–5.51 s) showed the highest variability between 212 individuals and populations. Temporal traits of the notes showed lower coefficients of 213 variation, and spectral measurements were highly stereotyped (Table 2, Figure 3). The lowest 214 and highest frequencies of the call were the most static properties. Overall (or among- 13 215 population) and within-population coefficients of variation were similar, indicating a 216 comparable magnitude in the variability of call traits regardless of grouping (Table 2). 217 218 Geographic variability in acoustic traits 219 Call properties differed between the sampling localities, although measurements 220 overlapped (Figure 4). Seven of the ten temperature-adjusted call traits differed among 221 populations, and five call traits differed between sides of the Amazon River (Table 3). 222 However, male body size also differed among the sampling sites (Nested ANOVA F = 3.68; p 223 = 0.001) and between river sides (Nested ANOVA F = 4.23; p = 0.042). Therefore, analyses of 224 variance were conducted with a new model, which included body size as covariate (nested 225 ANCOVA). As a result, we still found inter-population and between-river differences in most of 226 the call traits analysed. The differences were distributed among both temporal (related to the 227 notes and calls) and frequency traits (Table 3). The inclusion of body size as a co-variate 228 resulted in generally lower F statistics for the proposed models, indicating that interactive 229 effects between morphology and geography (sampling site) take place in the system. 230 Body-size differences were correlated with linear geographic distances (Mantel test r 231 = 0.669, p = 0.020) and not with the side of the Amazon River (Partial Mantel test r = -0.128, 232 p = 0.289). However, acoustic differences, after temperature and body size adjustments, did 233 not show correlations with linear geographic distances between population pairs (Mantel test r 234 = -0.029, p = 0.653) and river sides (Partial Mantel test r = -0.089, p = 0.449). 235 236 Discussion 237 Variability in acoustic sexual signal traits throughout the range of the Amazonian frog 238 Allobates paleovarzensis showed comparable magnitudes within and among populations, and 239 was independent of the degree of stereotypy of the acoustic characters. The relative 240 importance of a given call trait as a potential cue for discrimination can be assessed by 241 comparing the ratios between among- and within-population coefficients of variation (Pröhl et 242 al., 2007). Higher between- than within-population CV ratios were found in the polymorphic 243 Oophaga pumilio, a dendrobatoid species studied on a similar geographic scale (Pröhl et al., 14 244 2007). In a study comprising most of Brazilian Amazonia, the neotropical hylid frog 245 Dendropsophus leucophyllatus also showed a much lower within- than between-population 246 call variability, although this taxon may consist of a complex of cryptic species (Lougheed et 247 al., 2006). Therefore, our findings suggest that none of the call traits analysed is indicated as a 248 potential cue for social recognition or sexual selection between populations of A. 249 paleovarzensis. 250 Ecological and behavioural traits characteristic of the genus may be related to the 251 unexpectedly low overall-/within-population CV ratios observed in Allobates paleovarzensis. 252 Here, we hypothesise three non-exclusive causes: (1) Absence of local selective forces 253 causing signal differentiation. Its presence is expected in scenarios involving acoustic 254 character displacement (Lemmon, 2009). At all the localities visited in this study, no other 255 species of the ecologically similar A. trilineatus species group was found to occur with A. 256 paleovarzensis (Kaefer I. L., unpubl. data). Therefore, we believe that no ecological 257 interspecific selective pressures would be causing important call differentiation (i.e., higher 258 than between-populations) through the range of the species studied. However, it is important 259 to note that the composition of acoustically coactive sympatric frog species was not important 260 for intraspecific call differentiation in the congener A. femoralis (Amézquita et al., 2006). (2) 261 High call trait variability among males. Studies on A. femoralis suggest that between-male 262 call variation is pronounced and sufficient to allow statistical distinction of individual males 263 (Gasser et al., 2009). This individual signature in call traits could also be present in A. 264 paleovarzensis and lead to the high within-population levels of call variability observed. (3) 265 The use of visual cues to recognise and select mates. Interpopulation variation in calls may 266 have less importance in systems where visual cues are employed in mating interactions. It is 267 likely that A. femoralis uses visual components in mate recognition and choice (Luna et al., 268 2010; Montanarin et al., 2011), as was observed in the dendrobatoid Oophaga pumilio (Maan 269 & Cummings, 2009). In addition, the observation of multimodal communication (with the 270 visual component as the secondary channel) during agonistic interactions in the genus lends 271 further support to this hypothesis (Narins et al., 2003; Narins et al., 2005). Therefore, future 272 investigation involving playback experiments on intraspecific mate recognition in A. 273 paleovarzensis should provide us with elements to disentangle the candidate causes for the 274 observed pattern of similar within- and among-population call variability. 15 275 Only spectral properties of the advertisement call of Allobates paleovarzensis were 276 significantly affected by male body size. This effect is often observed in frogs, and is due to 277 the mass and resonance of the laryngeal sound-producing structures (McClelland et al., 1996), 278 although some species also show temporal properties influenced by male size (e.g., Zweifel, 279 1968; Castellano et al., 2002; Gasser et al., 2009). The high coefficients of determination 280 observed indicate that all static acoustic properties of the call are under strong morphological 281 constraint, and also that these call properties may be honest indicators of their control factors. 282 Therefore, we should expect that females might impose a stabilising selection on spectral 283 traits of A. paleovarzensis, although we cannot disentangle the selective forces driving body 284 size and spectral features in order to evaluate their relative importance in call evolution. 285 As a neuromuscular mechanism, the regulation of the temporal properties of calls for 286 anurans and other ectotherms is affected by environmental temperature (Prestwich, 1994). In 287 Allobates paleovarzensis, all temporal traits related to individual notes, but none related to 288 calls, were affected with this co-factor. This apparently lower sensitivity of calls to 289 temperature may be related to the lower stereotypy (i.e., higher coefficients of variation) of 290 call parameters in relation to those of the notes, making them subject to higher “confounding” 291 intrinsic variability. In fact, Gerhardt & Huber (2002, p. 15) reported a trend for shorter 292 components of the call to be more static, and our data support this observation. Additionally, 293 the social environment composed by co- and heterospecific acoustic signals at the time of the 294 recording may constitute a third source of variation that may affect call traits (Wells & 295 Schwartz, 2007). Therefore, given that call characters were not related to morphological or 296 temperature constraints, it seems likely that, in addition to intrinsic factors, their display may 297 be subject to social regulation. 298 Both spectral and temporal call traits varied among populations and between sides of 299 the Amazon River. Intraspecific geographic variation in acoustic signals is frequently 300 observed in Amazonian frogs, and highlights the neglected within-species diversity of frogs in 301 the biome, which is also home to a large number of undescribed species (Fouquet et al., 2007; 302 Angulo & Icochea, 2010). Differentiation in Amazonian anurans has been tested regarding 303 river barrier effects, and provided mixed results associated with geological, idiosyncratic river 304 characteristics (e.g., Gascon et al., 1996, 1998; Lougheed et al., 1999; Funk et al., 2007; 305 Simões et al., 2008; Tsuji-Nishikido et al., 2012). Among these characteristics, width has 306 been indicated as determining the effectiveness of a river barrier in promoting speciation in 16 307 Amazonian birds (Hayes & Sewlal, 2004) and mammals (Patton & da Silva, 1998). We are 308 not aware of any other study investigating the Amazon, the largest river of the biome, as a 309 promoter of phenotypic differentiation in frogs. Not all acoustic properties of A. 310 paleovarzensis showed congruent patterns of divergence, and both river and geographic 311 distance effects were not significant in the overall acoustic variability when summarising all 312 call traits in principal components. The Amazon River was found to be most effective as a 313 dispersal barrier for forest birds along its lower, wider portion, between its confluence with 314 the Negro River and its delta, where it meets the Atlantic Ocean (Hayes & Sewlal, 2004). 315 According to this proposal, our study area lies in a portion where the Amazon River is 316 expected to exert intermediate effects as a vicariant barrier. Future studies should be designed 317 to test the effect of the river on intraspecific differentiation along different portions of its 318 course. In relation to isolation-by-distance effects, a correlation between geographic and 319 acoustic distances would be indicative of neutral processes acting in phenotypic 320 differentiation. Therefore, our results indicate that local stabilising selective forces may be 321 important in call divergence. 322 There was a positive relation between body size and geographic distances among 323 populations, and as expected, spectral variables of the call followed a similar pattern of 324 variation. However, the different call properties showed divergent patterns of geographic 325 variability through the range of the species, as also observed for the dendrobatoid Oophaga 326 pumilio in Costa Rica and Panama (Pröhl et al., 2007). In addition, we could not associate the 327 divergent geographic patterns of call variation to the trait’s degree of stereotypy, as observed 328 for green toads from Central Asia (Castellano et al., 2000), which suggests that call variation 329 in some anuran species may not follow this pattern. 330 331 Acknowledgements 332 We are grateful to Ana C. Mello, Anelise Montanarin, Arthur O. Nascimento, Moisés S. 333 Melo, Priscila P. Correa, and Raimundo N. Amorim for fieldwork assistance; to Anderson S. 334 Bueno, Luciana K. Erdtmann, and Pedro I. Simões for valuable suggestions during this study; 335 to Janet W. Reid for the English revision; and to two anonymous reviewers whose 336 observations greatly improved this paper. This work is part of the PhD thesis of ILK. We also 337 thank the Brazilian Conselho Nacional de Desenvolvimento Científico e Tecnológico for 338 financial support (CNPq - CT Amazônia 575572/2008-6). CNPq provided a research 17 339 fellowship to ILK and a productivity grant to APL. This study was allowed by RAN340 ICMBio/IBAMA (licences 13777-2, 18516-1 and 20065-2). 341 342 References 343 Amézquita, A., Hödl, W., Lima, A.P., Castellanos, L., Erdtmann, L. & Araújo, M.C. (2006). 344 Masking interference and the evolution of the acoustic communication system in the 345 Amazonian dendrobatid frog Allobates femoralis. – Evolution 60: 1874-1887. 346 Amézquita, A., Lima, A.P., Jehle, R., Castellanos, L., Ramos, O., Crawford, A.J., Gasser, H. 347 & Hödl, W. (2009). Calls, colours, shape, and genes: a multi-trait approach to the study of 348 geographic variation in the Amazonian frog Allobates femoralis. – Biol. J. Linn. Soc. 98: 826349 838. 350 Anderson, M.J. & Legendre, P. (1999). An empirical comparison of permutation methods for 351 tests of partial regression coefficients in a linear model. – J. Stat. Comp. Simul. 62: 271-303. 352 Angulo, A. & Icochea, J. (2010). Cryptic species complexes, widespread species and 353 conservation: lessons from Amazonian frogs of the Leptodactylus marmoratus group (Anura: 354 Leptodactylidae). – Syst. Biodiv. 8: 357-370. 355 Bee, M.A., Cook, J.M., Love, E.K., O’Bryan, L.R., Pettitt, B.A., Schrode, K. & Vélez, A. 356 (2010). Assessing acoustic signal variability and the potential for sexual selection and social 357 recognition in Boreal Chorus Frogs (Pseudacris maculata). – Ethology 116: 564-576. 358 Behr, O., von Helversen, O., Heckel, G., Nagy, M., Voigt, C.C. & Mayer, F. (2007): 359 Territorial songs indicate male quality in the sac-winged bat Saccopteryx bilineata 360 (Chiroptera, Emballonuridae). – Behav. Ecol. 17: 810-817. 361 Bonnet, E. & Van de Peer, Y. (2002). ZT: a software tool for simple and partial Mantel tests. 362 – J. Stat. Soft. 7: 1-12. 363 Boul, K.E., Funk, W.C., Darst, C.R., Cannatella, D.C. & Ryan, M.J. (2007). Sexual selection 364 drives speciation in an Amazonian frog. – Proc. R. Soc. Lond., B. 274: 399-406. 18 365 Castellano, S., Cuatto, B., Rinella, R., Rosso, A. & Giacoma, C. (2002). The advertisement 366 call of the European treefrogs (Hyla arborea): a multilevel study of variation. – Ethology 108: 367 75-89. 368 Castellano, S. & Giacoma, C. (1998). Stabilizing and directional female choice for male calls 369 in the European green toads. – Anim. Behav. 56: 275-287. 370 Castellano, S., Giacoma, C. & Dujsebayeva, T. (2000). Morphometric and advertisement call 371 geographic variation in polyploid green toads. – Biol. J. Linn. Soc. 70: 341-360. 372 Chakraborty, M., Mangiamele, L.A. & Burmeister, S.S. (2010). Neural activity patterns in 373 response to interspecific and intraspecific variation in mating calls in the Túngara Frog. – 374 PLoS One 5: e12898. 375 Charif, R.A., Waack, A.M. & Strickman, L.M. (2008). Raven Pro 1.3 User’s Manual. – 376 Cornell Laboratory of Ornithology, Ithaca, New York. 377 Fischer, J., Kitchen, D.M., Seyfarth, R.M. & Cheney, D.L. (2004). Baboon loud calls 378 advertise male quality: acoustic features and their relation to rank, age, and exhaustion. – 379 Behav. Ecol. Sociobiol. 56: 140-148. 380 Fouquet, A., Gilles, A., Vences, M., Marty, C., Blanc, M. & Gemmell, N.J. (2007). 381 Underestimation of species richness in Neotropical frogs revealed by mtDNA analyses. – 382 PLoS One 32: e1109. 383 Funk, W.C., Caldwell, J.P., Peden, C.E., Padial, J.M., De la Riva, I. & Cannatella, D.C. 384 (2007). Tests of biogeographic hypotheses for diversification in the Amazonian Forest Frog, 385 Physalaemus petersi. – Mol. Phyl. Evol. 44: 825-837. 386 Gascon, C., Lougheed, S.C. & Bogart, J.P. (1996). Genetic and morphological variation in 387 Vanzolinius discodactylus: a test of the river hypothesis of speciation. – Biotropica 28: 376388 387. 389 Gascon, C., Lougheed, S.C. & Bogart, J.P. (1998). Patterns of genetic population 390 differentiation in four species of Amazonian frogs. A test of the Riverine Barrier Hypothesis. 391 – Biotropica 30: 104-119. 19 392 Gasser, H., Amézquita, A. & Hödl, W. (2009). Who is calling? Intraspecific call variation in 393 the aromobatid frog Allobates femoralis. – Ethology 115: 596-607. 394 Gerhardt, H.C. (1991). Female mate choice in treefrogs: static and dynamic acoustic criteria. 395 – Anim. Behav. 42: 615-635. 396 Gerhardt, H.C. & Huber, F. (2002). Acoustic Communication in Insects and Anurans: 397 Common Problems and Diverse Solutions. – The University of Chicago Press, Chicago. 398 Gomez, D., Richardson, C., Théry, M., Lengagne, T., Léna, J., Plénet, S. & Joly, P. (2011). 399 Multimodal signals in male European treefrog (Hyla arborea) and the influence of population 400 isolation on signal expression. – Biol. J. Linn. Soc. 103: 633-647. 401 Guerra, M.A. & Ron, S. (2008). Mate choice and courtship signal differentiation promotes 402 speciation in an Amazonian frog. – Behav. Ecol. 19: 1128-1135. 403 Hayes, F.E. & Sewlal, J.N. (2004). The Amazon River as a dispersal barrier to passerine 404 birds: effects of river width, habitat and taxonomy. – J. Biogeogr. 31: 1809-1818. 405 Heyer, R.W. & Reid, Y.R. (2003). Does advertisement call variation coincide with genetic 406 variation in the genetically diverse frog taxon currently known as Leptodactylus fuscus 407 (Amphibia: Leptodactylidae)? – An. Acad. Bras. Ciênc. 75: 39-54. 408 Lemmon, E.M. (2009). Diversification of conspecific signals in sympatry: geographic overlap 409 drives multidimensional reproductive character displacement in frogs. – Evolution 63: 1155410 1170. 411 Lima, A.P., Caldwell, J.P., Biavati, G. & Montanarin, A. (2010). A new species of Allobates 412 (Anura: Aromobatidae) from paleovárzea forest in Amazonas, Brazil. – Zootaxa 2337: 1-17. 413 Lougheed, S.C., Austin, J.D., Bogart, J.P, Boag, P.T. & Chek, A.A. (2006). Multi-character 414 perspectives on the evolution of intraspecific differentiation in a neotropical hylid frog. – 415 BMC Evol. Biol. 2006: 6-23. 416 Lougheed, S.C., Gascon, C., Jones, D.A., Bogart, J.P. & Boag, P.T. (1999). Ridges and rivers: 417 a test of competing hypotheses of Amazonian diversification using a dart-poison frog, 418 Epipedobates femoralis. – Proc. R. Soc. Lond., B. 266: 1829-1835. 20 419 Luna, A.G., Hödl, W. & Amézquita, A. (2010). Colour, size and movement as visual 420 subcomponents in multimodal communication by the frog Allobates femoralis. – Anim. 421 Behav. 79: 739-745. 422 Maan, M.E. & Cummings, M.E. (2009). Sexual dimorphism and directional sexual selection 423 on aposematic signals in a poison frog. – Proc. Natl. Acad. Sci. U. S. A. 106: 19072-19077. 424 McClelland, B.E., Wilczynski, W. & Ryan, M.J. (1996). Correlations between call 425 characteristics and morphology in male cricket frogs (Acris crepitans). – J. Exp. Biol. 199: 426 1907-1919. 427 Melendez, K.V. & Feng, A.S. (2010). Communication calls of little brown bats display 428 individual-specific characteristics. – J. Acoust. Soc. Am. 128: 919-923. 429 Montanarin, A., Kaefer, I.L. & Lima, A.P. (2011). Courtship and mating behaviour of the 430 brilliant-thighed frog Allobates femoralis from Central Amazonia: implications for the study 431 of a species complex. – Ethol. Ecol. Evol. 23: 141-150. 432 Morales, V.R. (2002 “2000”). Sistemática y biogeografia del grupo trilineatus (Amphibia, 433 Anura, Dendrobatidae, Colostethus), con descripción de once nuevas especies. – Publ. Asoc. 434 Amigos Doñana 13: 1-59. 435 Narins, P.M., Grabul, D.S., Soma, K.K., Gaucher, P. & Hödl, W. (2005). Cross-modal 436 integration in a dart-poison frog. – Proc. Natl. Acad. Sci. U. S. A. 102: 2425-2429. 437 Narins, P.M., Hödl, W. & Grabul, D.S. (2003). Bimodal signal requisite for agonistic 438 behavior in a dart poison frog, Epipedobates femoralis. – Proc. Natl. Acad. Sci. U. S. A. 100: 439 577-580. 440 Panhuis, T.M., Butlin, R., Zuk, M. & Tregenza, T. (2001). Sexual selection and speciation. – 441 Trends Ecol. Evol. 16: 364-371. 442 Patton, J.L. & da Silva, M.N.F. (1998). Rivers, refuges, and ridges: The geography of 443 speciation of Amazonian mammals. – In: Endless Forms: Species and Speciation (Howard, 444 D.J. & Berlocher, S.H., eds). Oxford University Press, Oxford, p. 202-213. 445 Prestwich, K.N. (1994). The energetics of acoustic signaling in anurans and insects. – Am. 446 Zool. 34: 625-643. 21 447 Pröhl, H., Hagemann, S., Karsch, J. & Höbel, G. (2007). Geographic variation in male sexual 448 signals in Strawberry Poison Frogs (Dendrobates pumilio). – Ethology 113: 825-837. 449 Rodríguez, A., Nuez, D. & Alonso, R. (2010). Intraspecific variation in the advertisement call 450 of the Cloud-Forest Frog Eleutherodactylus glamyrus (Anura: Eleutherodactylidae). – J. 451 Herpetol. 44: 457-466. 452 Ryan, M.J., Cocroft, R.B., & Wilczynski, W. (1990). The role of environmental selection in 453 intraspecific divergence of mate recognition signals in the cricket frog, Acris crepitans. – 454 Evolution 44: 1869-1872. 455 Ryan, M.J., Rand, A.S. & Weigt, L.A. (1996). Allozyme and advertisement call variation in 456 the tungara frog, Physalaemus pustulosus. – Evolution 50: 2435-2453. 457 Ryan, M.J. & Wilczynski, W. (1991). Evolution of intraspecific variation in the advertisement 458 call of a cricket frog (Acris crepitans, Hylidae). – Biol. J. Linn. Soc. 44: 249-271. 459 Simões, P.I., Lima, A.P., Magnusson, W.E., Hödl, W. & Amézquita, A. (2008). Acoustic and 460 morphological differentiation in the frog Allobates femoralis: relationships with the upper 461 Madeira River and other potential geological barriers. – Biotropica 40: 607-614. 462 Smith, M.J. & Hunter, D. (2005). Temporal and geographic variation in the advertisement call 463 of the Booroolong Frog (Litoria booroolongensis: Anura: Hylidae). – Ethology 111: 1103464 1115. 465 Tsuji-Nishikido, B.M., Kaefer, I.L., Freitas, F.C., Menin, M. & Lima, A.P. (2012). Significant 466 but not diagnostic: differentiation through morphology and calls in the Amazonian frogs 467 Allobates nidicola and A. masniger. – Herpetol. J. 22: 105-114. 468 Vences, M. & Wake, D.B. (2007). Speciation, species boundaries and phylogeography of 469 amphibians. – In: Amphibian Biology, Systematics (Heatwole, H. & Tyler, M.J., eds). Surrey 470 Beatty & Sons, Chipping Norton, p. 2613-2671. 471 Verburgt, L., Ferreira, M. & Ferguson, J.W.H. (2011). Male field cricket song reflects age, 472 allowing females to prefer young males. – Anim. Behav. 81: 19-29. 22 473 Wells, K.D. & Schwartz, J.J. (2007). The Behavioral Ecology of Anuran Communication. – 474 In: Hearing and Sound Communication in Amphibians (Narins, P.M., Feng, A.S., Fay, R.R. & 475 Popper, A.N., eds). Springer-Verlag, New York, p. 44-86. 476 Wilkinson, L. (2007). SYSTAT: the system for statistics. SYSTAT inc., Evanston. 477 Zweifel, R.G. (1968). Effects of temperature, body size, and hybridization on mating calls of 478 toads, Bufo a. americanus and Bufo woodhousii fowleri. – Copeia 1968: 269-285. 22 Table 1. The effect of temperature and body size on call traits in Allobates paleovarzensis (n = 123 males). Linear regression coefficients of determination (r2), F statistics and significance values (p) are presented. For body size (SVL), call traits were temperature-adjusted. Significant p values after Bonferroni adjustment are marked with asterisks. Temperature Body size Trait r2 F P r2 F P NR 0.27 45.14 <0.001* 0.00 0.87 0.353 ND 0.18 28.39 <0.001* 0.00 0.05 0.818 InI 0.21 33.08 <0.001* 0.00 1.56 0.213 CR 0.00 0.09 0.768 0.00 0.61 0.437 CD 0.00 0.91 0.341 0.02 4.11 0.045 IcI 0.00 0.00 0.981 0.00 0.19 0.664 MF 0.00 0.35 0.554 0.30 54.33 <0.001* LF 0.00 1.13 0.289 0.35 66.55 <0.001* HF 0.00 1.44 0.232 0.34 63.74 <0.001* NM 0.11 15.99 <0.001* 0.01 2.47 0.118 23 Table 2. Descriptive statistics of call traits based on overall acoustic parameter means calculated from mean values of 123 males recorded from ten localities throughout the range of Allobates paleovarzensis, representing natural variation between 23.1 and 28.0 oC air temperature. Mean within-individual, mean within-population and overall coefficients of variation (CVwi, CVwp, CVo, %) of ten traits of the advertisement call (adjusted to 25.4oC using linear regression) of Allobates paleovarzensis. Traits are classified by type based on mean within-individual CVs according to Gerhardt (1991). Trait Mean ± SD Range CVwi CVwp CVo CVo/CVwp Type NR 5.98 ± 1.37 3.56–10.14 11.1 ± 6.4 17.1 ± 4.2 19.7 1.15 Intermediate ND 0.04 ± 0.005 0.02–0.06 6.8 ± 3.6 10.5 ± 2.6 11.3 1.08 Intermediate InI 0.14 ± 0.04 0.07–0.29 16.0 ± 10.2 21.7 ± 3.8 26.3 1.21 Dynamic CR 0.27 ± 0.07 0.12–0.44 25.0 ± 14.0 23.8 ± 6.6 26.7 1.12 Dynamic CD 2.31 ± 1.03 0.75–8.73 31.8 ± 13.5 33.2 ± 8.7 44.8 1.35 Dynamic IcI 2.04 ± 0.75 1.02–5.51 28.1 ± 19.3 29.8 ± 9.7 34.3 1.15 Dynamic MF 4534.15 ± 192.47 4045.17–5092.62 1.2 ± 1.1 3.6 ± 0.7 4.1 1.12 Static LF 4147.06 ± 160.87 3645.14–4545.05 0.9 ± 0.7 3.1 ± 0.7 3.7 1.19 Static HF 4832.32 ± 187.83 4255.77–5316.92 1.0 ± 0.6 3.1 ± 0.6 3.6 1.16 Static NM 685.26 ± 104.98 462.00–968.51 6.5 ± 3.4 13.4 ± 2.4 14.3 1.07 Intermediate 24 Table 3. Nested analysis of (co-)variance of call traits (dependent variables) of Allobates paleovarzensis. Population (nested in river side) and river side are the independent categorical variables. In model b, size is used as the co-variate. Significant p values after Bonferroni adjustment are marked with asterisks. a. Nested ANOVA b. Nested ANCOVA (body size as co-variate) Trait F Pop P F River P F Pop P F River P NR 5.21 <0.001* 8.57 0.004* 5.04 <0.001* 8.09 0.005 ND 2.75 0.011 0.38 0.534 2.69 0.012 0.31 0.576 InI 7.41 <0.001* 2.86 0.093 7.09 <0.001* 2.68 0.104 CR 3.66 0.001* 16.79 <0.001* 3.75 0.001* 17.61 <0.001* CD 3.34 0.002* 0.00 0.966 3.23 0.003* 0.10 0.747 IcI 1.48 0.182 1.20 0.275 1.53 0.164 0.88 0.348 MF 1.96 0.066 19.20 <0.001* 0.62 0.731 14.30 <0.001* LF 3.40 0.002* 27.70 <0.001* 1.65 0.126 22.77 <0.001* HF 3.70 0.001* 16.80 <0.001* 2.52 0.019 11.97 <0.001* NM 3.17 0.004* 0.37 0.540 3.20 0.003* 0.78 0.37 25 Figure captions Figure 1. Geographic distribution of sampling locations in Central Amazonia. 1 = Careiro, 2 = Janauacá, 3 = Manaquiri, 4 = Hiléia, 5 = Novo Airão, 6 = Anamã, 7 = Anori, 8 = Codajás, 9 = Unini, 10 = Barcelos. Figure 2. (A) Oscillogram and (B) spectrogram of three multinote advertisement calls, and (C) power spectrum of a single note of a male Allobates paleovarzensis at its type locality in Careiro da Várzea, Brazil. Figure 3. Within-individual coefficients of variation (CVwi) for ten acoustic properties of the advertisement call of Allobates paleovarzensis. Grey boxes indicate the middle 50% of the data (interquartile range), and horizontal lines represent the median values. Dots represent outlying values. Vertical lines indicate the 10th and 90th percentiles. Figure 4. Variation in call traits (adjusted to 25.4 oC using linear regression) and body size across the range of Allobates paleovarzensis in Central Amazonia. Grey boxes indicate the middle 50% of the data (interquartile range), and horizontal lines represent the median values. Dots represent outlying values. Vertical lines indicate the 10th and 90th percentiles. 26 Fig. 1 27 Fig. 2 28 Fig. 3 29 Fig. 4 30 Capítulo 2 Kaefer, I.L.; Tsuji-Nishikido, B.M. & Lima, A.P. 2012. Beyond the river: underlying determinants of population acoustic signal variability in Amazonian direct-developing Allobates (Anura: Dendrobatoidea). Acta Ethologica, 15:187-194. 31 1 2 3 Beyond the river: underlying determinants of population acoustic signal variability in 4 Amazonian direct-developing Allobates (Anura: Dendrobatoidea) 5 Igor L. Kaefer1, Bruno M. Tsuji-Nishikido2 and Albertina P. Lima1 6 7 1 Programa de Pós-Graduação em Ecologia, Instituto Nacional de Pesquisas da Amazônia, Av. 8 Efigênio Sales, 2239, CP 478, CEP 69011–970, Manaus, Amazonas, Brazil. 9 2 Programa de Pós-Graduação em Diversidade Biológica, Universidade Federal do Amazonas, 10 Av. General Rodrigo Octavio, 3000, CEP 69077-000, Manaus, Amazonas, Brazil. 11 Corresponding author: Igor L. Kaefer. E-mail address: [email protected] 12 13 Abstract The multidimensional nature of animal signals makes acoustic traits potentially 14 subject to different determinants. The Amazonian frogs Allobates nidicola and Allobates 15 masniger have an allopatric distribution, occurring along the left and right sides of the 16 Madeira River, respectively. These are two sister, phenotypically similar species whose eggs 17 are deposited and develop entirely in a terrestrial nest. In this study, we analyzed 2,000 18 advertisement calls recorded from ten localities across Central Amazonia, in order to 19 understand the role of determinants of acoustic signal variability at the population and species 20 levels. We assessed, through nested analyses of variance, the differentiation of six characters 21 of this sexual signal among populations and between interfluves. Moreover, we measured the 22 degree of variability and the extent of temperature- and body size-induced plasticity in call 23 traits. We also tested for isolation-by-distance effects in phenotypic differentiation through 24 Mantel tests. Coefficients of variation were higher among than within populations for all call 25 measurements. Spectral call properties were more distinctive than temporal traits among 26 populations and species. Advertisement call traits showed strong temperature-induced 27 plasticity (e.g., 45% of the variation in note duration). In contrast, the effects of body size 28 were restricted to frequency-related characters. The river barrier effect was significant among 29 all the acoustic variables analyzed even after controlling for male body size. Geography 30 (sampling locality) and body size also jointly affected call variability. No correlation between 32 31 geographical and acoustic distances among populations was observed, suggesting that local 32 stabilizing selective pressures have an important role in the evolution of call differentiation. 33 Keywords advertisement call, geographical variation, isolation by distance, river barrier 34 effect, sexual signal, signal plasticity. 35 36 Introduction 37 Each of the properties of an acoustic signal typically shows a characteristic range of variation 38 at different levels of analysis (e.g., within populations, among populations, between species) 39 (Castellano et al. 2002). Moreover, it is expected that the multidimensional nature of animal 40 signals will make acoustic traits potentially subject to different determinants (Gerhardt 1991; 41 Gerhardt and Huber 2002). By decomposing acoustic signals into their properties, we might 42 be able to assess the agents responsible for call variation and to account for their effects 43 depending on the question asked in a particular study system. Additionally, by measuring the 44 degree of variability of quantitatively variable traits of frog calls in different levels, we can 45 evaluate their potential for sexual selection and social recognition, and also understand the 46 roles of particular acoustic characters in microevolutionary processes (e.g., Friedl and Klump 47 2002; Gasser et al. 2009; Bee et al. 2010). 48 Call traits show temperature- and body size-induced plasticity (Prestwich 1994). 49 Environmental temperature is often regarded as a cofactor in acoustic studies, and it is 50 necessary to account for its effect, especially for ectotherms, in order to obtain more accurate 51 measures of call differentiation when comparing individuals, populations, or species. The 52 extent to which temperature affects call traits varies considerably in anurans, and tropical 53 species, which undergo more subtle temperature variations, are less studied than are species in 54 temperate regions (see Gerhardt and Huber 2002; Wells 2007). Another cofactor often 55 controlled in comparative acoustic studies is male body size, which is known to affect 56 predominantly frequency-related characters of the call (Ryan 1988). 57 Geography plays an important role in the evolution and differentiation of acoustic 58 signals (Wycherley et al. 2002), but its effect may not be equivalent for all the properties of 59 the call. It has been suggested that spatial variation takes place according to the degree of 60 variability of a given acoustic character, with static properties exhibiting a more pronounced 33 61 variation among populations (Castellano et al. 2000). Moreover, relationships between 62 geography and covariables such as body size might produce interactive effects between 63 determinants of call differentiation (Pröhl et al. 2007; Kaefer and Lima 2012). 64 Biogeographical studies in Amazonia are revealing the effects of river barriers on call 65 divergence for different frog species, both in overall acoustic variability (Simões et al. 2008; 66 Tsuji-Nishikido et al. 2012) and in specific properties of the call (Kaefer and Lima 2012). 67 However, isolation by distance is also recognized as an important mechanism in phenotypic 68 divergence in Amazonian frogs (e.g., Lougheed et al. 2006; Pröhl et al. 2006; Amézquita et 69 al. 2009). Because isolation-by-distance effects are inferred through correlations between 70 phenotypic and distance measures, they can be masked by the presence of river barriers, 71 which might lead to acoustic distances that are disproportionally greater than those produced 72 by linear distance. Therefore, testing for the occurrence of isolation-by-distance effects might 73 provide clues to inform the selection/drift debate regarding the evolutionary forces driving 74 differentiation. 75 In an earlier study, we demonstrated the usefulness of advertisement calls in 76 discriminating between the similar Amazonian species Allobates nidicola Caldwell and Lima 77 2003 and A. masniger Morales 2002 (Tsuji-Nishikido et al. 2012). This study indicated that 78 the lower portion of the Madeira River is an important biogeographical barrier involved in 79 phenotypic differentiation, and suggested that A. nidicola and A. masniger have an allopatric 80 distribution, occurring along the left and right sides of the river, respectively. Molecular 81 genetic data confirm this proposed distribution and provide evidence that they constitute sister 82 species (Kaefer et al. unpublished data). A. nidicola and A. masniger are two diurnal, ground83 dwelling species whose eggs are deposited and develop entirely in a terrestrial nest, and their 84 adults share all the available morphological diagnostic characters (Morales 2002; Caldwell 85 and Lima 2003) (Fig. 1). Tsuji-Nishikido et al. (2012), as in most taxonomic studies, limited 86 themselves to summarizing the call measurements by applying principal components 87 analyses, thus overlooking the individual contribution of each call trait to the differentiation 88 between populations or species. Moreover, possible effects of isolation by distance were not 89 addressed. Therefore, in this study we characterized the underlying determinants of 90 population acoustic variability, by exploring this study system through the following 91 questions: (1) What is the degree of variability of each of the call traits among populations 92 and between species? (2) To what extent do temperature and body size induce call trait 93 variability? (3) Is the river barrier effect significant among all the different acoustic variables? 34 94 (4) Does isolation by distance play a significant role in call divergence, beyond the known 95 effects of the river as a barrier? 96 97 Material and methods 98 We analyzed a total of 2,000 calls from 100 males (20 calls per male) recorded at ten sites 99 located in the Brazilian states of Amazonas and Pará, including the type localities of A. 100 nidicola and A. masniger (Fig. 2). Ten individuals were sampled in each locality, which were 101 regarded as populations in this study. Data were collected between December 2009 and 102 March 2010, during the rainy season, which encompasses the breeding periods of both 103 species. Recordings were made using a Marantz PMD660 digital recorder (44.1 kHz sampling 104 rate; 16-bit resolution) and an AKG 568 EB directional microphone positioned 1-1.5 m in 105 front of the calling male. After each recording, we measured environmental (air) temperature 106 with a digital thermometer and captured the frog to measure body size (snout-vent length, 107 SVL, in millimeters) with a digital caliper. The advertisement call of both species is 108 composed of a single note repeated rhythmically over time, thus forming sequences of notes 109 lasting from 2 to 5 min. A comparative graphical representation of the advertisement call of 110 both species can be found in Tsuji-Nishikido et al. (2012). Call measurements were taken on 111 Raven 1.2 (Window: Blackman; Discrete Fourier Transform: 1,024 samples; 3 dB filter 112 bandwidth: 80.0 Hz) (Charif et al. 2004). The measurement used to represent each individual 113 in the statistical analysis was the mean value of the parameters obtained from 20 notes and 114 their respective intervals. We considered the following traits: note duration (ND, seconds); 115 note modulation (NM, Hertz); interval between notes (IN, seconds); lowest frequency (LF, 116 Hertz); highest frequency (HF, Hertz) and peak frequency (PF, Hertz). Call measurements are 117 illustrated in Fig. 3. 118 Our data set was checked for normality prior to the parametric statistical analyses. In 119 order to assess the effect of temperature on call properties, we performed linear regression 120 analyses for all populations combined. For all within- and among-population analyses, we 121 used temperature-adjusted acoustic data by calculating overall regression coefficients (b) for 122 all call parameters at the mean temperature of 25.9 oC (SD = 1.77, range = 23.1–31.2) through 123 the equation Y adj = y – (b*tcalling site) + (b*tmean) where Y adj = adjusted value of call 124 measurement, b = regression coefficient, and t = temperature. The relationship between 35 125 temperature-adjusted acoustic traits and male body size was also assessed by linear regression 126 analyses. The adjusted r2 values obtained were used to estimate the percentage of variation 127 explained by the independent variable. 128 Coefficients of variation (CV = SD*100/mean) for each call trait were calculated 129 within populations (CVwp), within species (CVws), and over all individuals analyzed (CVo). 130 Our objectives were to evaluate the degree of variability of each call measurement and to 131 determine which call parameters might be most distinctive among populations and between 132 species. Within-population CVs are based on the mean and SD over all call-recorded 133 individuals in each population (n = 10). We obtained the mean within-population CV (CVwp) 134 by averaging all within-population CVs. Within-species CVs are based on the mean and SD 135 over all call-recorded individuals in each side of the river. We obtained the mean within136 species CV (CVws) by averaging the two within-species CVs. Overall CVs (CVo) were 137 obtained from the grand mean and SD over all populations and were initially calculated for 138 individuals (n = 100) of all populations combined (Pröhl et al., 2007). 139 To compare the difference in calls (with temperature-adjusted traits) between opposite 140 sides of the Madeira River (species) and among populations, we conducted nested analyses of 141 variance (ANOVA), with populations nested in river sides. Two different models for each call 142 trait were used: (a) a nested ANOVA without cofactor and (b) a nested ANOVA with size as 143 cofactor (nested ANCOVA). These approaches allowed us to compare the F statistics 144 obtained through the inclusion of body size as a cofactor in the analyses: if an interaction 145 takes place between cofactors and populations, a decline in F statistics is expected due to 146 worsening of the model. 147 The correlation between linear geographical/body size distances among sampling sites 148 and the respective acoustic distances was evaluated through Mantel tests (Smouse et al. 149 1986). A body size distance matrix was obtained through the calculation of Euclidean 150 distances between mean SVL measurements for each sampling locality. The acoustic distance 151 matrix was calculated by conducting a principal components analysis from the data set of 152 temperature-adjusted call measurements. We obtained pairwise Euclidean distances between 153 all possible pairs of populations by using the scores of the mean acoustic measurements in the 154 first and second principal components (which together explained 71.4% of the total acoustic 155 variation). Mantel tests were conducted in ZT (Bonnet and van de Peer 2002), using 36 156 permutation of the null models (Anderson and Legendre 1999), and applying 10,000 157 randomizations. 158 159 Results 160 The coefficients of variation, which indicate the degree of variability of the call traits, differed 161 when within-population, within-species, and overall variability were considered. Considering 162 the within-species and overall perspectives, most of the call traits analyzed showed 163 intermediate coefficients of variation. Regardless of grouping, spectral acoustic traits were the 164 less variable (Fig. 4). They also showed the highest CVo/CVwp and CVo/CVws ratios, which 165 suggests that spectral characters are more distinctive than temporal traits among populations 166 and species (Table 1). 167 Environmental temperature significantly affected all call measurements except note 168 modulation. Body size induced variation of all spectral properties of the call (Table 2). 169 Acoustic traits differed statistically among the populations. The F statistics of spectral trait 170 differentiation decreased when body size was included as a covariate. However, among171 population differentiation was still significant for the highest frequency after this procedure. 172 The lower Madeira River showed a strong effect on acoustic differentiation for all acoustic 173 traits considered, even when the effects of male body size were controlled (Table 3). 174 When all localities sampled were considered, the Mantel test revealed a strong 175 association between the advertisement call and body size differences between population 176 pairs. Geographical distances were not correlated with acoustic or body size distances. When 177 the Mantel tests were restricted to the left and right sides of the Madeira River (species), no 178 correlation regarding the traits analyzed was found between population pairs (Table 4). 179 180 Discussion 181 What is the degree of variability of each of the call traits among populations and 182 between species? 37 183 Coefficients of variation were higher in the overall perspective and reached its lower 184 values within populations for all call traits analyzed (i.e., CVwp < CVws < CVo). Our results 185 showed that spectral call properties were more distinctive than temporal traits among 186 populations and species, with higher distinctiveness at the among-population level compared 187 to the between-species level (i.e., CVo/CVws < CVo/CVwp). Therefore, our understanding of 188 the effect of river barriers in speciation processes would be greatly benefited by future 189 approaches that evaluate the degree of stereotypy of call traits at the within-individual level, 190 as well as the role of spectral properties of the advertisement call in mate recognition and 191 choice between individuals from opposite river sides. 192 To what extent do temperature and body size induce call trait variability? 193 In spite of the small range of variation in air temperature characteristic of tropical 194 environments (23.1 to 31.2 oC in this study), there was strong temperature-induced plasticity 195 in call traits. Note duration had 45% of its variation induced by this variable. Spectral 196 characters of the call were also significantly affected by air temperature, although to a lesser 197 extent (6–7%). Significant effects on frequency-related call measurements were also observed 198 in Oophaga pumilio throughout its range (Pröhl et al. 2007), but not in Allobates femoralis in 199 a single locality (Gasser et al. 2009). As ectotherms, frogs likely have spectral call traits that 200 are indirectly subject to the strong physiological constraints imposed by the environmental 201 and consequently the body temperature (Prestwich 1994). 202 The effects of body size on the advertisement call of the individuals were restricted to 203 frequency-related characters, as is most commonly reported for anurans (Gerhardt and Huber 204 2002). Body size-induced plasticity was pronounced in these measurements, ranging from 41 205 to 43%. Because of the invariant nature of these call properties, it was suggested that spectral 206 characters of the advertisement call may be important indicators of male size (e.g., Bee et al. 207 2000; Smith and Hunter 2005). Body-size effects are frequently removed in comparative 208 population analyses. However, as ontogenetic differences may be successfully controlled by 209 this procedure, natural variability in adult SVL, which might be important and even 210 diagnostic in comparative studies, is overlooked by applying body-size corrections (e.g., 211 Castellano et al. 2000; Pröhl et al. 2007). Therefore, despite the difficulties in differentiating 212 between the two sources of variability related to body size, the description of these 213 relationships is important to the investigation of the morphological constraints of acoustic 214 variation. Interestingly, the properties of note modulation and internote interval were little 38 215 affected by temperature and the size of the frogs. This suggests that additional underlying 216 factors, such as physiological state, body condition, and social environment (Castellano and 217 Giacoma 1998; Tárano 2001; Wells and Schwartz 2007), might be responsible for the residual 218 (i.e., unexplained) variability in the call traits analyzed. 219 Is the river barrier effect significant among all the different acoustic variables? 220 The river barrier effect was significant among all the different acoustic variables 221 analyzed, even after the effects of male body size were controlled. The inclusion of SVL as a 222 covariate lowered the F statistics obtained in the analysis of spectral measurements, 223 suggesting an interactive effect between geography and body size on call differentiation. Our 224 results in the nested analysis of variance complement the findings of Tsuji-Nishikido et al. 225 (2012) by addressing the role of the lower Madeira River in affecting each of the call traits. 226 The effect of Amazonian rivers has been observed in a range of organisms, evidencing its 227 important role in evolutionary processes in the biome (e.g., Ayres and Clutton-Brock 1992; 228 Peres et al. 1996; Hayes and Sewlal 2004; Simões 2010). In contrast, the among-population 229 differentiation regarding call traits was severely reduced after we accounted for body-size 230 effects and was restricted to the properties of note duration, interval between notes, and 231 highest frequency. Thus, the overall ANOVA results show a predominant effect of the 232 Madeira River in addition to the among-population differences in the acoustic signals studied. 233 Does isolation by distance play a significant role in call divergence beyond the known 234 effects of the river as a barrier? 235 No correlation between geographic distance (a potential predictor of both degree of 236 environmental similarity and gene flow) and acoustic (sexual signal) dissimilarities was 237 revealed by the Mantel tests, and the observed pattern held when the analyses were restricted 238 to each side of the river. This apparent absence of isolation-by-distance effects was also 239 observed for call differentiation in Allobates paleovarzensis in Central Amazonia (Kaefer and 240 Lima 2012), but not for the neotropical frogs Dendropsophus leucophyllatus and A. femoralis 241 on wider geographical scales (Lougheed et al. 2006; Amézquita et al. 2009). This suggests 242 that local stabilizing selective pressures may be more important than neutral effects, such as 243 genetic drift, in the evolution of call differentiation in the A. nidicola/masniger system. 244 Regarding the two geographical factors presumed to be driving phenotypic differentiation 245 among the populations studied – river and linear geographical distance – our results give 39 246 further support to the relationship between sexual signal differentiation and the Madeira River 247 as a barrier, as indicated by Tsuji-Nishikido et al. (2012). A molecular assessment of 248 population variability, together with data on phenotypic variation, will seek to assess the 249 relationship between genetic and phenotypic divergence between A. nidicola and A. masniger, 250 a study system that shows the potential to elucidate various aspects regarding the initial stages 251 of allopatric speciation. 252 Acknowledgements We are grateful to Francisco C. de Freitas for fieldwork assistance, 253 Marcelo Menin for the loan of recording equipment, Pedro Ivo Simões for valuable 254 suggestions during this study, and to Janet W. Reid and anonymous reviewers for comments 255 on earlier versions of this manuscript. This work is part of the Ph.D. thesis of ILK. The 256 Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, Brazil) provided 257 financial support (CT Amazônia 553997-2006-8 and 575572-2008-6), a research fellowship 258 to ILK, and a productivity grant to APL. The Coordenação de Aperfeiçoamento de Pessoal de 259 Nível Superior (CAPES, Brazil) provided a research fellowship to BMT. This study was 260 approved by RAN-ICMBio/IBAMA (permit # 21950-1). 261 262 References 263 Amézquita A, Lima AP, Jehle R, Castellanos L, Ramos O, Crawford AJ, Gasser H, Hödl W 264 (2009) Calls, colours, shape, and genes: a multi-trait approach to the study of geographic 265 variation in the Amazonian frog Allobates femoralis. Biol J Linn Soc 98:826-838 266 Anderson MJ, Legendre P (1999) An empirical comparison of permutation methods for tests 267 of partial regression coefficients in a linear model. J Stat Comp Simul 62:271-303 268 Ayres JM, Clutton-Brock TH (1992) River boundaries and species range size in Amazonian 269 primates. Am Nat 140:531-537 270 Bee MA, Cook JM, Love EK, O’Bryan LR, Pettitt BA, Schrode K, Vélez A (2010) Assessing 271 acoustic signal variability and the potential for sexual selection and social recognition in 272 boreal chorus frogs (Pseudacris maculata). Ethology 116:564-576 40 273 Bee MA, Perrill SA, Owen PC (2000) Male green frogs lower the pitch of acoustic signals in 274 defense of territories: a possible dishonest signal of size? Behav Ecol 11:169-177. 275 Bonnet E, Van de Peer Y (2002) ZT: a software tool for simple and partial Mantel tests. J Stat 276 Soft 7:1-12 277 Caldwell JP, Lima AP (2003) A new Amazonian species of Colostethus (Anura: 278 Dendrobatidae) with a nidicolous tadpole. Herpetologica 59:219-234 279 Castellano S, Cuatto B, Rinella R, Rosso A, Giacoma C (2002) The advertisement call of the 280 European treefrogs (Hyla arborea): a multilevel study of variation. Ethology 108:75-89 281 Castellano S, Giacoma C (1998) Stabilizing and directional female choice for male calls in 282 the European green toads. Anim Behav 56:275-287 283 Castellano S, Giacoma C, Dujsebayeva T (2000) Morphometric and advertisement call 284 geographic variation in polyploid green toads. Biol J Linn Soc 70:341-360 285 Charif RA, Clark CW, Fristrup KM (2004) Raven 1.2 user’s manual. Cornell Laboratory of 286 Ornithology, Ithaca 287 Friedl TWP, Klump GM (2002) The vocal behaviour of male European treefrogs (Hyla 288 arborea): implications for inter- and intrasexual selection. Behaviour 139:113-136 289 Gasser H, Amézquita A, Hödl W (2009) Who is calling? Intraspecific call variation in the 290 aromobatid frog Allobates femoralis. Ethology 115:596-607 291 Gerhardt HC (1991) Female mate choice in treefrogs: static and dynamic acoustic criteria. 292 Anim Behav 42:615-635 293 Gerhardt HC, Huber F (2002) Acoustic communication in insects and anurans: common 294 problems and diverse solutions. The University of Chicago Press, Chicago 295 Hayes FE, Sewlal JN (2004) The Amazon River as a dispersal barrier to passerine birds: 296 effects of river width, habitat and taxonomy. J Biogeogr 31:1809-1818 297 Kaefer IL, Lima AP (2012) Sexual signals of the Amazonian frog Allobates paleovarzensis: 298 geographic variation and stereotypy of acoustic traits. Behaviour 149:15-33 41 299 Lougheed SC, Austin JD, Bogart JP, Boag PT, Chek AA (2006) Multi-character perspectives 300 on the evolution of intraspecific differentiation in a neotropical hylid frog. BMC Evol 301 Biol 2006:6-23 302 Morales VR (2002) Sistematica y biogeografía del grupo trilineatus (Amphibia, Anura, 303 Dendrobatidae, Colostethus), con descripción de once nuevas especies. Publ Asoc 304 Amigos Doñana 13:1-59 305 Peres CA, Patton JL, da Silva MNF (1996) Riverine barriers and gene flow in Amazonian 306 saddle-back tamarins. Folia Primatol 67:113-124 307 Prestwich KN (1994) The energetics of acoustic signaling in anurans and insects. Am Zool 308 34:625-643 309 Pröhl H, Hagemann S, Karsch J, Höbel G (2007) Geographic variation in male sexual signals 310 in Strawberry Poison Frogs (Dendrobates pumilio). Ethology 113:825-837 311 Pröhl H, Koshy RA, Mueller U, Rand AS, Ryan MJ (2006) Geographic variation of genetic 312 and behavioral traits in northern and southern túngara frogs. Evolution 60:1669-1679 313 Ryan, MJ (1988) Energy, calling, and selection. Am Zool 28:885-898 314 Simões PI (2010) Diversificação do complexo Allobates femoralis (Anura, Dendrobatidae) 315 em florestas da Amazônia brasileira: desvendando padrões atuais e históricos. Ph.D. 316 Thesis, Instituto Nacional de Pesquisas da Amazônia 317 Simões PI, Lima AP, Magnusson WE, Hödl W, Amézquita A (2008) Acoustic and 318 morphological differentiation in the frog Allobates femoralis: Relationships with the 319 upper Madeira River and other potential geological barriers. Biotropica 40:607-614 320 Smith MJ, Hunter D (2005) Temporal and geographic variation in the advertisement call of 321 the booroolong frog (Litoria booroolongensis: Anura: Hylidae). Ethology 111:1103-1115 322 Smouse PE, Long JC, Sokal RR (1986) Multiple regression and correlational extensions of 323 the Mantel test of matrix correspondence. Syst Zool 35:627-632 324 Tárano Z (2001) Variation in male advertisement calls in the neotropical frog Physalaemus 325 enesefae. Copeia 2001:1064-1072 42 326 Tsuji-Nishikido BM, Kaefer IL, Freitas FC, Menin M, Lima AP (2012) Significant but not 327 diagnostic: Differentiation through morphology and calls in the Amazonian frogs 328 Allobates nidicola and A. masniger. Herpetol J 22:105-114 329 Wells KD (2007) The ecology and behavior of amphibians. The University of Chicago Press, 330 Chicago 331 Wells KD, Schwartz JJ (2007) The behavioral ecology of anuran communication. In: Narins 332 PM, Feng AS, Fay RR, Popper AN (eds) Hearing and sound communication in 333 amphibians. Springer, New York, pp 44-86 334 Wycherley J, Doran S, Beebee TJC (2002) Male advertisement call characters as 335 phylogeographical indicators in European water frogs. Biol J Linn Soc 77:355-365 43 Tables Table 1 Descriptive statistics of the call traits analyzed, based on mean individual call properties of 100 individual males, adjusted to 25.9 °C using linear regression. Within-population (CVwp), within-species (CVws), and overall (CVo) coefficients of variation (%, given as mean ± SD) of six traits of the advertisement call of Allobates nidicola and A. masniger. Trait Mean ± SD; A. nidicola Mean ± SD; A. masniger CVwp CVws CVo CVo/CVwp CVo/CVws ND (s) 0.048 ± 0.005 0.050 ± 0.005 7.1 ± 2.3 9.4 ± 0.6 9.6 1.36 1.01 NM (Hz) 354.7 ± 46.7 396.9 ± 51.8 12.5 ± 2.6 13.1 ± 0.1 14.2 1.14 1.08 IN (s) 0.290 ± 0.048 0.339 ± 0.086 16.2 ± 5.4 20.9 ± 6.0 23.3 1.44 1.11 LF (Hz) 4048.9 ± 255.6 4299.3 ± 231.1 4.5 ± 1.9 5.8 ± 0.6 6.5 1.45 1.12 HF (Hz) 4403.9 ± 253.7 4699.4 ± 207.2 3.8 ± 1.3 5.0 ± 0.9 6.0 1.58 1.18 PF (Hz) 4222.5 ± 244.3 4497.6 ± 209.7 4.0 ± 1.4 5.2 ± 0.8 6.1 1.50 1.16 44 Table 2 The effect of temperature and body size on call traits of the individuals studied (n = 100 males). Adjusted linear regression coefficients of determination (r2), F statistics and significance values (p) are presented. For body size, statistics were calculated using temperature-adjusted call traits. Significant p values after Bonferroni adjustment are marked with asterisks Temperature Body size Trait r2 F P r2 F P Note duration 0.45 82.00 <0.001* -0.003 0.62 0.562 Note modulation -0.003 0.69 0.589 -0.01 0.00 0.985 Internote interval 0.07 8.62 0.004* 0.03 3.60 0.057 Lowest frequency 0.07 8.15 0.005* 0.41 69.64 <0.001* Highest frequency 0.06 7.84 0.006* 0.43 74.59 <0.001* Peak frequency 0.07 8.61 0.004* 0.42 73.10 <0.001* 45 Table 3 Nested analysis of (co-)variance of call traits (dependent variables) of A. nidicola and A. masniger. Population (nested in river side) and river side (species) are the independent categorical variables. In model b, size is used as covariate. Significant p values after Bonferroni adjustment are marked with asterisks a. Nested ANOVA b. Nested ANCOVA (size) Trait F Pop P F River P F Pop P F River P ND 8.62 <0.001* 7.72 0.006* 8.67 <0.001* 7.49 0.007* NM 1.34 0.232 18.79 <0.001* 1.30 0.249 24.08 <0.001* IN 8.60 <0.001* 19.99 <0.001* 8.53 <0.001* 10.22 0.001* LF 6.09 <0.001* 37.35 <0.001* 2.14 0.039 9.66 0.002* HF 8.02 <0.001* 64.01 <0.001* 3.58 0.001* 22.41 <0.001* PF 6.58 <0.001* 53.13 <0.001* 2.55 0.014 16.92 <0.001* 46 Table 4 Mantel tests evaluating correlations between acoustic (DAco), body size (DSVL), and geographical (DGeo) distances of Allobates sampled at ten localities in Central Amazonia. Because of river effects, tests were also restricted to the left and right sides of the Madeira River (five localities each). Significant values after Bonferroni correction are marked with an asterisk. Model r p DAco X DGeo -0.15 0.312 DAco X DSVL 0.642 0.001* DSVL X DGeo -0.193 0.205 DAco X DGeo -0.484 0.108 DAco X DSVL 0.409 0.191 DSVL X DGeo 0.152 0.366 DAco X DGeo -0.355 0.166 DAco X DSVL 0.45 0.091 DSVL X DGeo -0.419 0.058 All localities Left side (A. nidicola) Right side (A. masniger) 47 Figures Fig. 1 A. nidicola (top) and A. masniger (bottom) males vocalizing at their respective type localities in Brazilian Amazonia Fig. 2 Sampling sites visited in this study: 1 Km 12 of Autazes Road, type locality of A. nidicola; 2 BR-319, km 260; 3 PPBio Manaquiri; 4 BR-319 Tupana; 5 Vila Gomes; 6 Borba; 7 Novo Aripuanã; 8 Road to Apuí; 9 Jacareacanga; 10 PARNA da Amazônia, type locality of A. masniger Fig. 3 A graphical representation of the call measurements adopted in this study. Oscillogram (a) and spectrogram (b) of a call sequence of a topotypic A. masniger male (SVL = 18.5 mm; INPA-H 28196), recorded at air temperature of 23.7 °C. Power spectrum (c) of a single call of this individual. Highest and lowest frequencies were measured at 20 dB below the peak intensity (horizontal dotted line), the value at which the signal energy could still be clearly distinguished from background noise. Note modulation was calculated as the difference between the highest and the lowest frequencies of the call Fig. 4 Within-population coefficients of variation (CV, in percent) for six acoustic properties of the advertisement call of the individuals studied. Gray boxes indicate the middle 50% of the data (interquartile range), and horizontal lines represent the median values. Vertical lines indicate the 10th and 90th percentiles. Note that the lowest coefficients of variation are related to spectral traits of the call 48 Fig. 1 49 Fig. 2 50 Fig. 3 51 Fig. 4 52 Capítulo 3 Kaefer, I.L.; Tsuji-Nishikido, B.M.; Mota, E.P.; Farias, I.P. & Lima, A.P. The early stages of speciation in Amazonian forest frogs: phenotypic conservatism despite strong genetic structure. Submetido à Evolutionary Biology 53 1 2 3 The early stages of speciation in Amazonian forest frogs: phenotypic conservatism 4 despite strong genetic structure 5 6 Igor L Kaefer 1*, Bruno M Tsuji-Nishikido 2, Edvaldo P Mota 2, Izeni P Farias 2, Albertina P 7 Lima 1 8 9 1 Programa de Pós-Graduação em Ecologia, Instituto Nacional de Pesquisas da Amazônia, CP 10 478, 69011-970, Manaus, Amazonas, Brazil 11 2 Instituto de Ciências Biológicas, Universidade Federal do Amazonas, 69077-000, Manaus, 12 Amazonas, Brazil 13 14 *Corresponding author 15 Email: [email protected] 16 Telephone: 55 (92) 81018404 17 Fax: 55 (92) 36431909 18 19 20 21 22 23 24 54 25 Abstract 26 Phylogeographic perspectives incorporating multiple classes of characters, especially those 27 relating to sexual signals, are promising for the elucidation of recent evolutionary 28 mechanisms driving speciation. Here, forest frogs were used as a model system to access 29 distinct stages in the process of evolutionary differentiation. We studied 280 individuals 30 assigned to three species: Allobates paleovarzensis, A. nidicola and A. masniger. Samples 31 were collected at 20 localities arranged in two study systems, along the middle Amazon and 32 the lower Madeira Rivers, in Central Amazonia. Mantel tests, analyses of molecular variance, 33 and the spatial distribution of haplogroups indicated that the distribution of genetic variability, 34 as inferred from a mitochondrial DNA marker, was determined by a combination of isolation35 by-distance effects and the transposition of large Amazonian rivers. These two factors had 36 contrasting relative influences in each of the study systems, which also differed regarding the 37 estimated time of the major cladogenetic events. Pronounced population genetic structure was 38 observed. However, multivariate discriminant function analyses showed that the phenotypic 39 (morphological and acoustic) divergence was loosely related with genetic differentiation and 40 did not successfully predict assignment of individuals to genetic groups. The observed 41 distribution of genetic variability showed the important role of genetic drift in the 42 diversification of the mitochondrial marker studied. The phenotypic conservatism observed 43 among populations was surprising when considering the high genetic structuring observed, 44 and indicates a prevailing role of stabilizing selective forces in the process of sexual signal 45 and morphological differentiation. 46 Keywords 47 genetic drift, integrative phylogeography, isolation by distance, mitochondrial DNA, river 48 barrier, sexual signals. 49 50 Introduction 51 Historically isolated populations are primer candidates for the study of the origin of 52 differences in genetic and phenotypic characters because they would have been potentially 53 exposed to long term genetic drift or divergent selection pressures, without the homogenizing 54 influences of gene flow (Avise 2000; Nosil et al. 2009). Therefore, it is expected that spatially 55 55 structured populations, which experience restricted demographic connectivity, might show 56 concordant divergences among different sets of characters. Nevertheless, decoupled patterns 57 of population differentiation can be observed, indicating that either stabilizing or directional 58 selection pressures acting differentially on specific characteristics may lead to distinct patterns 59 of the distribution of variation among populations and have important implications for the 60 speciation process (Santos et al. 2003; Adams et al. 2009; Lemmon 2009). 61 Recent theoretical and analytical advances enabled by the integration of distinct 62 disciplines have led to a better understanding of patterns and mechanisms of differentiation 63 among populations of diverse organisms (Crisci et al. 2003; Santini et al. 2012). Nonetheless, 64 the relationship between genotypic and phenotypic differentiation in the early stages of 65 speciation still remains poorly understood, directing questions to ecological and historical 66 mechanisms that promote or maintain variation at this level of the biological hierarchy (Diniz67 Filho et al. 2008). Important advancements may be obtained by investigating the geographic 68 pattern in which genetic variability is distributed among populations, as well as the order 69 according to which ecological and phenotypic differentiation appears (West-Eberhard 1989; 70 Dall 1997; Panhuis et al. 2001). 71 The extent and shape of the spatial structure of intraspecific genealogical relationships 72 are typically unknown for natural populations, especially in megadiverse tropical regions, 73 where not even the species boundaries are clearly defined (e.g., Groeneveld et al. 2009; 74 Vieites et al. 2009; Hubert et al. 2012). Beyond the challenge of delimiting species, a better 75 comprehension of the speciation process requires the study of geographical factors which may 76 underpin the origin and diversification of lineages within species, such as vicariance and 77 isolation by distance (Hutchison and Templeton 1999; Zink et al. 2000). In Amazonia, species 78 distribution patterns (e.g., Wallace 1852; Haffer 1969; Ayres and Clutton-Brock 1992) and 79 fine-scale genealogical structure assessments (e.g., Cohn-Haft 2000; Funk et al. 2007; 80 Solomon et al. 2008; Amézquita et al. 2009) have inferred the effect of these two historical 81 factors in terrestrial organisms, although the relative contribution of each of them varied 82 greatly, perhaps due to idiosyncratic particularities of the study systems. Still, no clear 83 geographical pattern or palaeoenvironmental force has emerged as a general explanation for 84 the diversification of most of the animal groups (Antonelli et al. 2010; Ribas et al. 2011). 85 Phylogeographic perspectives involving more than one class of character are 86 particularly promising for the elucidation of recent evolutionary mechanisms driving 56 87 differentiation, given that the history inferred from DNA sequences may provide a historical 88 demographic context to evaluate divergence in various phenotypic attributes (Wiens 2008, 89 Campbell et al. 2010; Guillot et al. 2012). These appraisals might be conducted with any kind 90 of genetically based marker, not necessarily of molecular origin (Avise 2004). This is because 91 many phenotypic attributes carry phylogenetic and even phylogeographic signals (Wycherley 92 et al. 2002; Erdtmann and Amézquita 2009; Goicochea et al. 2009) that might, together with 93 gene trees, reciprocally illuminate the history of populations and the mechanisms responsible 94 for evolutionary divergence. Integrative, multicharacter approaches are particularly important 95 for our knowledge on tropical organisms, which urgently need detection of significant 96 evolutionary units for conservation, assessment of cryptic diversity, and determination of 97 species limits (Moritz et al. 2000; Dayrat 2005; Wiens 2007; Crawford et al. 2010). 98 Phylogeographic structure is commonly observed among species of frogs, and is often 99 credited to their low vagility when compared to other vertebrates (Avise 2000; Vences and 100 Wake 2007). Frogs are suitable for population-level multi-character evolutionary approaches 101 because of their high abundance, which allows dense sampling along their distributional 102 ranges, and their conspicuous advertisement calls, that can be used as population markers 103 (Wycherley et al. 2002; Lougheed et al. 2006). Given that anurans use advertisement calls for 104 sexual recognition and mate choice, differences in these signals dramatically increase the 105 evolutionary potential for speciation due to the possibility of pre-zygotic isolation (Panhuis et 106 al. 2001; Pröhl et al. 2006; Boul et al. 2007). 107 Allobates (Anura, Dendrobatoidea, Aromobatidae) frogs were used here as a model 108 system to access distinct stages of the evolutionary differentiation process through the 109 examination of morphological, acoustic and genetic characters. We studied individuals 110 assigned to three nominal species: Allobates paleovarzensis Lima et al. (2010), A. nidicola 111 (Caldwell and Lima 2003) and A. masniger (Morales 2002). Sampling was performed on 20 112 localities arranged in two distinct geographic study systems. The choice of these systems was 113 based on studies of phenotypic variability that suggested that these populations may have 114 experienced initial, but distinct stages in the process of allopatric speciation (Kaefer and Lima 115 2012; Kaefer et al. 2012; Tsuji-Nishikido et al. 2012). This is the first population genetic 116 assessment of the three aforementioned species. This study focuses on the using of gene trees 117 and phenotypic data in order to disentangle most recent events that led to the spatial 118 organization of the multi-trait variation currently observed. 57 119 We expected that that a neutral, stochastic process plays a central role in the 120 emergence of evolutionary differentiation which may lead to the formation of new species. 121 Thus, we tested the premise that the distribution of genetic variability is determined by a 122 combination of the effects of isolation by distance (i.e., geographic distance among individual 123 samples) and the transposition of vicariant barriers (Amazonian rivers). Our null hypothesis 124 was that the phenotypic differentiation is related to the genetic pattern observed. Therefore, 125 we expected that phenotypic divergence would be correlated with genetic differentiation and 126 not incongruent with the groupings indicated by molecular analyses. In order to test this 127 hypothesis, we accessed the genealogical relationship among individual samples through 128 phylogeographic analyses, including parameters related to historical demography inferred 129 from a neutrally evolving DNA marker. We also evaluated the extent of population 130 differentiation in advertisement calls and morphology by their correlation with geographic 131 and genetic distances, as well as the roles of both acoustic and morphological traits in 132 predicting the assignment of individuals to genetic groups. 133 134 Methods 135 Study area 136 This study was conducted in two adjacent areas of the Central Amazon lowlands, in 137 the Brazilian states of Amazonas and Pará, on a region characterized by wide tracts of tropical 138 rainforests. These areas are crossed by the two major rivers of the biome: the Amazon, at its 139 middle course (Fig. 1A), and the Madeira, at its lower portion (Fig. 2A). The Madeira River is 140 the main southern tributary of the Amazon, which represents the largest river system on Earth. 141 The white, sediment-rich water channels of these rivers are characterized by intense 142 sedimentary dynamics in the Central Amazon, permitting the formation of marginal lakes, 143 islands and channels and by extensive floodplains covered with seasonally flooded forests 144 (várzeas) on its riverbanks (Sioli 1984; Irion and Kalliola 2010). The region also has 145 extensive plateaus harboring non-flooded rainforest (terra firme), as well as transitional 146 forests, called palaeo-várzeas, that are ancient floodplains of the Amazon River and its 147 tributaries (Irion and Kalliola 2010). 58 148 The paleovarzensis system 149 The Amazonian palaeo-várzea forests constitute the habitat for Allobates 150 paleovarzensis, a diurnal aromobatid frog which deposits its eggs in the forest floor and 151 whose tadpole development must be completed in water bodies. This is a recently described 152 species which distribution encompasses its type and adjacent localities south of the Amazon 153 River, and also localities to the north, reaching the municipality of Barcelos, at the right 154 margin of the Negro River (Kaefer and Lima 2012). Our sampling design includes all known 155 areas of occurrence of this species, including its type locality (Fig. 1A). 156 The nidicola-masniger system 157 Allobates nidicola and Allobates masniger are diurnal leaf-litter frogs that occur in 158 terra firme forests of Central Amazonia, having been described without reference to each 159 other. Recently, it was found that these species share all the available diagnostic characters, 160 including the direct, terrestrial development of endotrophic larvae into froglets (Tsuji161 Nishikido et al. 2012). Nevertheless, this study detected significant differences in morphology 162 and calls between populations from both type localities and from opposite sides of the lower 163 Madeira River, the main biogeographic barrier in the studied area. This report also suggests 164 that they constitute allopatrically distributed sister species. In order to better understand the 165 speciation process occurring in this study system, we analyze these two species jointly, 166 according to the sample design presented in Tsuji-Nishikido et al. (2012) (Fig. 2A). 167 Data collection 168 Data were collected at 20 sites (10 per study system), which will be referred to as 169 populations hereafter. The fieldwork procedures employed for the acquisition of the 170 phenotypic data analyzed here are fully described in recent studies which evaluated the 171 taxonomic identity and the variation in phenotypic characters of individuals belonging to both 172 study systems (Kaefer and Lima 2012; Kaefer et al. 2012; Tsuji-Nishikido et al. 2012). The 173 collection of specimens constituted the first assessment of the taxonomic identity of 174 individuals through the distributional ranges of the three species involved. The number of 175 individuals analyzed for each data set, and the geographic coordinates of the sample sites, are 176 provided in Table 1. Muscular tissue samples from each individual were extracted and 177 preserved in 95 percent ethanol prior to fixation of voucher specimens in formol. Tissue 178 samples were housed at the Coleção de Tecidos de Genética Animal of the Universidade 59 179 Federal do Amazonas (CTGA – UFAM), Manaus, Brazil. Studied individuals were deposited 180 in the Herpetological Collection of the Instituto Nacional de Pesquisas da Amazônia, in 181 Manaus (INPA-H). 182 Acoustic data. Call parameters from individuals of the paleovarzensis (10 183 calls/individual) and nidicola-masniger (20 calls/individual) systems were obtained from 184 recordings with Raven Pro 1.3 (Charif et al. 2008) and Raven 1.2 (Charif et al. 2004), 185 respectively. Temporal and spectral call traits analysed for each species are described on 186 Online Resource 1 and Online Resource 2. These measurements were averaged in order to 187 represent each of the populations in the statistical analyses. The call structure of each of the 188 species, including graphic representations, and the detailed description of call traits analyzed 189 are provided elsewhere (Kaefer and Lima 2012; Kaefer et al. 2012; Tsuji-Nishikido et al. 190 2012). Summary statistics of the call measurements are given on Online Resource 1 and 191 Online Resource 2. 192 Morphological data. External morphometric measurements (in mm) were taken from 193 the left side of the specimens, with a stereoscopic microscope with a graduated ocular lens 194 (precision 0.10 mm) and a digital caliper (to the nearest 0.01 mm). All individuals were 195 analyzed according to the twenty-three measurements presented and described in detail in 196 Online Resource 3. Among these measurements, snout-vent length (SVL) was used 197 exclusively as a covariate in analyses of phenotypic differentiation. Descriptive statistics for 198 the morphometric measurements in each sampling locality are given in Online Resource 3 and 199 Online Resource 4. 200 Molecular data. Genomic DNA was extracted using a standard cetyl trimethyl 201 ammonium bromide (CTAB) protocol. For the amplification of 16S ribosomal RNA gene, the 202 primers 16Sar and 16Sbr were used according to Palumbi (1996). This gene is the most 203 commonly used marker for amphibian systematics (Fouquet et al. 2007), including studies on 204 the superfamily Dendrobatoidea (Vences et al. 2000; Grant et al. 2006; Santos et al. 2009). It 205 shows high rates of amplification success (Vences et al. 2005), having satisfactory 206 performance at the intraspecific level, as well as among closely related species (Lougheed et 207 al. 2006; Vences et al. 2005; Simões 2010). 208 Polymerase Chain Reaction (PCR) amplifications were performed in 15 µL reaction 209 volumes containing 5.8 µL ddH2O, 1.5 µL of 25mM MgCl2, 1.5 µL of 10X amplification 210 buffer (75mM Tris HCl, 50 mM KCl, 20 mM (NH4)2SO4), 1.5 µL of a 2 µM solution of each 60 211 primer, 1.5 µL of 10 mM dNTPs (2.5mM each dNTP), 0.7 µL of Taq DNA polymerase 2.5 212 U/µL (Biotools, Spain) and 1 µL of template DNA (about 40 ng/ µL). The temperature profile 213 consisted of 1) preheating at 90 °C for 60 s, 2) denaturation at 92 °C for 60 s, 3) primer 214 annealing at 50 °C for 50 s, 4) extension at 72 °C for 90 s, and 5) a final extension at 72 °C 215 for 5 min. Steps 2-4 were repeated 35 times. Amplicons were purified with exonuclease and 216 alkaline phosphatase (Fermentas Life Sciences, Canada), and sequenced using forward 217 primers and ABI BigDye Terminator Cycle Sequencing Kit protocols, as indicated by the 218 manufacturer. The resulting single-stranded products were read in an ABI 3130xl automatic 219 capillary sequencer. The DNA sequences had their homologous regions aligned in BioEdit 220 (Hall 1999), where a preliminary alignment was generated by the ClustalW algorithm 221 (Thompson et al. 1994), and were finally checked and corrected by eye in comparison with 222 the original chromatograms. The final alignments were 527 bp long for Allobates 223 paleovarzensis, and 539 bp long for A. nidicola and A. masniger. Newly determined 224 sequences were deposited in GenBank (Online Resource 5). 225 Data analysis 226 Population analyses. The genealogical relationships among individual samples and 227 populations were estimated through haplotype networks using statistical parsimony 228 (Templeton et al. 1992) in TCS 1.21 (Clement et al. 2000). Gaps were considered as 5th 229 character state, and the connection limit was fixed at 23 steps to allow association among all 230 sampled haplotypes. We estimated the relative partitioning of genetic variation (within 231 populations, among populations, and between riversides) through standard molecular analysis 232 of variance (AMOVA; Excoffier et al. 1992), in Arlequin version 3.5 (Excoffier and Lisher 233 2010). The relative genetic structuring among sampling localities was estimated through the 234 calculation of distance-based fixation indices (FST; Wright 1951), equivalent to NST of Lynch 235 and Crease (1992). The statistical significance of FST values was obtained after 1023 236 haplotype permutations in Arlequin. Genetic differentiation among sampling localities was 237 estimated via Kimura 2-parameter corrected genetic distances (Kimura 1980) in MEGA 238 version 5 (Tamura et al. 2011). 239 The most likely number of genetic clusters formed by the sampled mitochondrial DNA 240 sequences was inferred through Bayesian analysis of population structure, in BAPS version 5 241 (Corander et al. 2008). Based on nucleotide frequencies, the model attempts to create k groups 242 of individuals, such that those allocated in the same group resemble each other genetically as 61 243 much as possible. The upper limit to the number of clusters was set from one to the number of 244 sample localities on each study system. The log-likelihood values of the best models were 245 used to select the most probable cluster arrangement. 246 Within each of the genetic clusters inferred by the Bayesian approach, we calculated 247 haplotype and nucleotide diversities, as well as the number of segregating sites, and 248 performed the neutrality tests Tajima’s D, Fu’s Fs, and Ramos-Onsis & Rozas’ R2 (Tajima 249 1989; Fu 1997; Ramos-Onsins and Rozas 2002). The statistical significance of these tests was 250 estimated through coalescent simulations with 10000 replicates in DnaSP 5.10 (Librado and 251 Rozas 2009). In addition to the neutrality tests, past demographic expansion events were 252 inferred from the distribution of the observed pairwise nucleotide site differences (mismatch 253 distributions) in each cluster, as calculated in DnaSP 5.10. We used coalescent simulation 254 procedures in order to test the validity of the estimated expansion models. The parameters of 255 these models were re-estimated in order to obtain their empirical confidence intervals, and the 256 empirical distribution of the sum of squared deviations between the observed and the 257 expected mismatch values (SSD; Schneider and Excoffier 1999) and the raggedness index 258 (Hri; Harpending 1994). These calculations were obtained through 10000 parametric 259 bootstrap replicates in Arlequin. 260 Divergence time estimation. Divergence times were estimated using BEAST 1.6.2 261 (Bayesian Evolutionary Analysis Sampling Trees; Drummond and Rambaut 2007). We 262 analyzed only unique haplotypes recovered from the population analyses conducted in 263 Arlequin. An available 16S sequence of Allobates talamancae (GenBank AY843577) was 264 considered as an external group for both systems. In addition, 16S sequences of A. trilineatus 265 (GenBank DQ502118) and A. brunneus (GenBank EU342522) were used as outgroups for the 266 paleovarzensis and nidicola-masniger systems, respectively. As priors (calibration points), we 267 used average means and confidence intervals of the most recent common ancestors of both the 268 external groups and the studied systems obtained from Santos et al. (2009). Models of 269 nucleotide substitution were estimated for the reduced mitochondrial DNA datasets in MEGA 270 version 5 (Tamura et al. 2011). For the two study systems, we applied an uncorrelated 271 lognormal relaxed clock model, a lognormal prior distribution, a randomly generated starting 272 tree, and length of chain of 10 million generations with samples taken every 10 thousand 273 generations, and discarding 10% of the trees as burn-in, resulting in 900 sampled trees in the 274 Markov Chain Monte Carlo (MCMC). The stationarity of the posterior distributions for all 275 model parameters, including medians and 95% Highest Posterior Density intervals (HPD) of 62 276 the nodes was checked on Tracer 1.5 (Rambaut and Drummond 2007). From the MCMC 277 output, we generated the final consensus tree using Tree Annotator 1.6.2 (Drummond and 278 Rambaut 2007). 279 Phenotypic differentiation. In order to test whether the genetic clusters assigned via 280 Bayesian analysis of population structure were phenotypically distinguishable, we used 281 discriminant function analyses (DFA) with acoustic and morphometric traits as predictor 282 variables. The discriminant function was used to calculate the probability of classification of 283 each individual collected in its genetic cluster (grouping variable) by a Jackknifed 284 classification matrix. This approach included all DNA-sequenced individuals with available 285 measurements regarding acoustic and/or morphometric traits. Acoustic data were both 286 temperature and temperature-SVL adjusted through linear regression. In order to account for 287 the effect of body size on morphometric measurements, we used 12 morphological ratios as 288 predictor variables (Verdade and Rodrigues 2007): LL/SVL, HAND3/SVL, FL/SVL, 289 HW/HL, EN/HL, EL/HL, TYM/HL, TYM/EL, IN/HW, HAND1/HAND2, HAND2/HAND3, 290 HAND4/HAND1 (see Online Resource 3 for acronyms). 291 Geographic correlates of phenotypic and genetic variation. We tested correlations 292 between the linear geographic distances (measured in km) between sampling sites and genetic 293 and phenotypic distances between the populations of each of the study systems by applying 294 Mantel tests on distance matrices derived from genetic, acoustic and morphological data sets 295 (Mantel 1967). In addition, we conducted partial Mantel tests to evaluate correlations between 296 genetic/phenotypic distances among populations of each study system and their separation by 297 the river channel while controlling for effects of geographic distance between sampling sites 298 (Smouse et al. 1986; Telles et al. 2001). We constructed a binary correspondence matrix, 299 designating the values “0” and “1” for localities within the same and between opposite sides 300 of the river, respectively. 301 Average genetic distances (16S) between sampling localities were calculated 302 according to the Kimura 2-parameter (K2P) model in MEGA version 5. Acoustic distance 303 matrices were obtained from the dataset of temperature-adjusted call measurements by 304 calculating pairwise Euclidean distances between all possible pairs of populations by using 305 the scores of the mean acoustic measurements on first and second components (which 306 together explained 79.0% of the total acoustic variation in A. paleovarzensis and 79.3% in A. 307 nidicola/masniger) produced by a principal components analysis (PCA). The principal 63 308 components were obtained from the arithmetic means of the call traits among all sampled 309 individuals for each of the localities, and were used in order to reduce the number of 310 independent phenotypic variables. Morphological distance matrices were generated from the 311 external measurements (except SVL) through the same procedure. The first two components 312 explained 64.3% of the total morphological variation in A. paleovarzensis and 71.1% in A. 313 nidicola/masniger. As body size was correlated with the scores of the first principal 2 314 component of both acoustic (linear regression r = 0.51, F = 9.39, p = 0.017 in A. 315 paleovarzensis, and r2 = 0.72, F = 23.78, p = 0.001 in A. nidicola/masniger) and 2 316 morphological data sets (linear regression r = 0.83, F = 43.65, p < 0.001 in A. 317 paleovarzensis, and r2 = 0.95, F = 174.09, p < 0.001 in A. nidicola/masniger), we regressed 318 them against corresponding mean SVL measurements for each population, and used the 319 residuals as new, size-independent acoustic variables from which new Euclidean distances 320 were calculated. Body size adjustments are useful for controlling possible ontogenetic effects 321 that might affect the outcome of comparative results. However, they can mask important call 322 differentiation related to body size dissimilarities among populations and species, which were 323 reported in the studied systems (Kaefer and Lima 2012; Kaefer et al. 2012). Therefore, we 324 also calculated acoustic distance matrices without applying SVL corrections. Mantel tests 325 were conducted in ZT (Bonnet and Van der Peer 2002) using permutation of the null models 326 (Anderson and Legendre 1999), and applying 10 000 randomizations. 327 328 Results 329 Population analyses 330 We obtained 276 16S rDNA sequences (Table 1), corresponding to 55 different 331 haplotypes (Online Resource 5). The distribution of haplotypes among the sampled localities 332 and the large number of mutations separating the genealogical groups showed a pronounced 333 genetic structure among populations, with restricted haplotype sharing across riverbanks in A. 334 paleovarzensis, while absent in the nidicola-masniger system (Fig. 1B and 2B). The AMOVA 335 indicated that a small fraction of the overall genetic variation can be found within sampling 336 localities, being concentrated among populations in the paleovarzensis system (60.37%) and 337 between opposite riverbanks in the nidicola-masniger system (59.74%) (Table 2). The FST 338 values obtained for both study systems revealed an overall high and significant population 64 339 structuring (Table 3). Genetic distances were more pronounced between populations of the 340 nidicola-masniger system (reaching 5.2%) than those observed in A. paleovarzensis (reaching 341 1.7%), thus lending further support to the interspecific nature of the former. 342 Bayesian analysis of genetic structure supported the partition of the individuals of A. 343 paleovarzensis into four genetic clusters (log ML value = -525.13; probability = 0.99), while 344 seven groups were estimated for the nidicola-masniger system (log ML value = -1182.18; 345 probability = 1.00). Co-occurrence of genetic clusters within sampling localities was nearly 346 absent. The distribution of sampling localities exerted a clear effect on the determination of 347 the genetic groupings, especially in A. nidicola and A. masniger, where six populations 348 corresponded to completely segregated clusters (Fig. 3). The three neutrality tests indicated 349 historical demographic changes in two clusters composed by individuals from almost the 350 entire range of A. paleovarzensis. In contrast, in the nidicola-masniger system the cluster 351 composed by the locality Road to Apuí was the only one to show significant departures from 352 historical demographic equilibrium according to the three neutrality tests, including the R2, 353 which is the most appropriate for small samples (Ramos-Onsins and Rozas 2002) (Table 4). 354 The tests of demographic growth based on the sum of squared deviation (SSD) between 355 observed and expected mismatch distributions, and Harpending’s raggedness index (Hri) did 356 not reject the demographic expansion hypothesis, giving support for the significant values 357 observed in the neutrality tests (Table 5, Online Resource 6). 358 Divergence time estimation 359 The Bayesian evolutionary analysis indicated the time of the first divergence between 360 the clades of A. paleovarzensis occupying the north and south sides of the middle Amazon 361 River to be on Pleistocene (median = 1.24 million years (mya); HPD = 0.54, 2.23). The 362 median time of the most recent common ancestor for the species was estimated at 2.29 mya 363 (HPD = 1.23, 3.81), and most of the intraspecific cladogenetic events in A. paleovarzensis are 364 indicated as having occurred during the last 1 million years (Fig. 4A). The split of the 365 nidicola-masniger system between the left and right sides of the lower Madeira River and the 366 time of the most recent common ancestor for this study system did not have occurred after 367 Early Pliocene (HPD = 6.88, 15.33; median = 10.65 mya;), with the subsequent cladogenetic 368 events evenly distributed from the time of this divergence to the present (Fig. 4B). 65 369 Phenotypic differentiation 370 Discriminant function analyses using genetic clusters as grouping variables was able 371 to correctly classify only 50.0% of the individuals of Allobates paleovarzensis according to 372 temperature-adjusted acoustic characters, and 43.0% of the individuals with both temperature 373 and body size-adjusted traits. Based on morphological characters, the percentage of correct 374 membership assignment of A. paleovarzensis was even poorer: 38.0% (Online Resource 7). 375 Individuals of A. nidicola and A. masniger were more successfully classified than those of A. 376 paleovarzensis in their respective genetic clusters based on acoustic characters adjusted for 377 temperature (59.0%), and both temperature and body size (48.0%). Morphological characters 378 also performed worse than acoustic traits within the nidicola-masniger system (41.0% of 379 individuals correctly classified; Online Resource 8). 380 Geographic correlates of phenotypic and genetic variation 381 The Mantel tests showed a significantly positive correlation between geographic and 382 genetic distances in both study systems, indicating a strong role of stochastic processes in the 383 process of evolutionary diversification. In contrast, correlations between phenotypic (acoustic 384 and morphological) and geographic distances were not significant (Table 6). We did not 385 observe signal of the Amazon River as a barrier promoting differentiation in A. 386 paleovarzensis in genetic or phenotypic distances. On the other hand, the Partial Mantel tests 387 revealed a role of the lower Madeira River as a vicariant barrier on the nidicola-masniger 388 system. Significant correlations were found between side of river and both overall genetic 389 differentiation and uncorrected acoustic distances. 390 391 Discussion 392 Pronounced population genetic structure 393 The high levels of genetic structure at the mitochondrial level found among 394 populations, especially in A. nidicola and A. masniger, are notable even considering earlier 395 reports on this subject, which indicate that panmictic scenarios are the exception among 396 amphibian populations (Crawford 2003; Zeisset and Beebee 2008). The main life history 397 characteristics pointed out to promote and maintain high genetic structure levels in anurans, 398 such as small body size and strong site attachment (Vences and Wake 2007) are widely 66 399 reported among Allobates (Pröhl 2005; Kaefer et al. in press). In addition, the studied species 400 are reported to be territorial (Lima et al. 2010; Caldwell and Lima 2003), and to exhibit direct 401 larval development in the nidicola-masniger system (Tsuji-Nishikido et al. 2012). Given that 402 water bodies are believed to constitute the main extraterritorial resource for dendrobatoid 403 species (Pröhl 2005), the absence of tadpole transport behavior may further restrict the use of 404 space of adult frogs, as well as concentrate offspring close to parental individuals, 405 contributing to the limited gene flow among lineages. The effects of this presumed high 406 phylopatry in A. nidicola and A. masniger may also be observed from the results of the 407 analysis of molecular variance, where a low within-population molecular variability was 408 observed for this study system (only 2.02 % of the total variation). 409 The configuration of genetic clusters recovered from the Bayesian analyses showed 410 significant genetic structuring over short geographic distances, resembling the pattern 411 observed in the haplotype networks. The star-shaped network of A. paleovarzensis, together 412 with the results of the neutrality tests, indicated that this species experienced recent 413 demographic changes across most of its range. Such patterns were less pronounced in the 414 molecular data of A. nidicola and A. masniger. Although the population growth tests did not 415 reject the population expansion hypothesis, the highly structured haplotype network and the 416 non-significant results in neutrality tests in most of the genetic clusters did not support past 417 demographic expansion events within A. nidicola and A. masniger. A mixed result in 418 neutrality tests was also observed through the range of A. femoralis along the Madeira River, 419 and might be related to demographic expansion of determinate populations which experienced 420 dispersal events, probably allowed by historical river course vagility (Simões 2010). 421 Spatio-temporal contexts of the inferred vicariant events 422 Both study systems exhibited signatures in their genetic structures that are congruent 423 with the transposition of large Amazonian rivers. The effectiveness of the lower Madeira 424 River as a barrier to gene flow in Allobates was observed to be higher than that of the middle 425 Amazon River. This observation is supported by the (1) absence of haplotype sharing among 426 sides of the Madeira River (which was restricted to different banks of the Amazon); (2) 427 accountability for most (60.37%) of the genetic variability in A. nidicola and A. masniger (the 428 river barrier was the second major factor explaining molecular variation within A. 429 paleovarzensis); (3) effect on overall variability of acoustic characters (significant differences 430 between interfluves were restricted to determinate call traits in A. paleovarzensis; (Kaefer and 67 431 Lima 2012); (4) Existence of reciprocally monophyletic clades on opposite sides of the 432 Madeira river (were paraphyletic between sides of the Amazon). This last characteristic 433 satisfies even the most stringent predictions of the classic river barrier hypothesis (Gascon et 434 al. 1996; Gascon et al. 1998), indicating the completeness of the speciation process in the 435 nidicola-masniger system according to both the evolutionary (sensu Wiley 1978) and 436 phylogenetic (Rosen 1978) species concepts. Under neutrality, the usual chronology of 437 evolutionary events leading towards speciation begins with the stage of polyphyly in a gene 438 tree, with successive paraphyly stages before the eventual achievement of the status of 439 reciprocal monophyly (de Queiroz 2007). According to this model, the monophyly observed 440 in the gene tree of the nidicola-masniger system is expected to arise from old geographic 441 barriers, in which diverging populations have experienced the time necessary to achieve such 442 status. Nevertheless, non-monophyly (both paraphyly and polyphyly) can be observed in 443 many of the currently recognized species (Funk and Omland 2003; McKay and Zink 2009). 444 Idiosyncratic characteristics of the sampling design and the history of the studied 445 species can be related the differential roles of the rivers as barriers that have been reported 446 (Colwell 2000). In this study, these system-specific characteristics may include: (1) Different 447 sampling designs: while the effect of the Madeira River as a barrier was accessed in its lower 448 portion, where it is believed to play a significant role as barrier to gene flow (Simões 2010), 449 A. paleovarzensis occurs along the middle section of the Amazon River, where it was 450 observed to exert intermediate effects as a vicariant barrier (Hayes and Sewlal 2004). (2) 451 Distinct study models: although belonging to the same genus, aspects of the natural history of 452 the studied species, such as dispersal abilities, might be responsible for distinct 453 microevolutionary arrangements (see discussion on genetic structure above). (3) 454 Asynchronous divergence times: besides the large credibility intervals estimated by the 455 Bayesian analyses, the median dates for the split of clades inhabiting opposite sides of the 456 river in both study systems indicates that these events took place in different geological 457 epochs. The cladogenic event separating haplotypes of A. paleovarzensis inhabiting the north 458 and south of the Amazon River probably occurred during the Pleistocene, being younger that 459 the divergence time estimated between A. nidicola and A. masniger, which might have not 460 occurred after the early Pliocene. According to these findings, we infer that A. paleovarzensis 461 might not have undergone the level and duration of genetic isolation required for population 462 differentiation to cross the threshold necessary to met the criteria of speciation under 463 monophyly- and phenotype-based species concepts (Coyne and Orr 2004). 68 464 Our timetree suggests that the lineages of A. paleovarzensis occupying opposite 465 riverbanks of the middle Amazon River diverged during the late phase of its formation, when 466 its fan experienced an expressive increase in sedimentation rates and high vagility in the 467 course of its drainage (Rossetti et al. 2005; Figueiredo et al. 2009). Therefore, although the 468 Amazon River initiated as a transcontinental watercourse between 11.8 and 11.3 Mya ago, we 469 believe that sedimentation-induced channel course dynamics, which are evidenced by 470 Quaternary palaeochannels and meander relicts (Costa et al. 2001), allowed dispersal events 471 across its middle section. The late Miocene split between A. nidicola and A. masniger at the 472 portion of the Madeira River where it meets the Amazon seems to match the onset of the 473 latter as a channel running from the Andes to the Atlantic Ocean (Hoorn 1994; Figueiredo et 474 al. 2009). The estimated age of this evolutionary split is in accordance with studies that 475 indicated that most of the speciation events in Amazonia largely predated the Pleistocene 476 (Hoorn et al. 2010; Ribas et al. 2011). For amphibians, the origin of most of the current 477 Amazonian species is estimated to have occurred during late Miocene (Santos et al. 2009; 478 Funk et al. 2012; this study), with estimates reaching the late Oligocene (Elmer et al. 2007; 479 Fouquet et al. 2012). 480 The present appraisal of the effects of rivers as barriers must be interpreted cautiously 481 as they represent byproducts of a multi-character approach conducted in species with 482 restricted distributions. Investigation on this subject should involve species (or species 483 groups) that can be found throughout the course of a river to be assessed, thus allowing for a 484 multiple paired sampling design (e.g., Silva and Patton 1998; Simões 2010). In addition, 485 nuclear markers should be considered in future genetic studies in order to access their 486 potential in disentangle chronological aspects on the history of the studied populations (Zhang 487 and Hewitt 2003). In spite of the limitations of our data to provide accurate details on the 488 temporal and spatial mechanisms by which these vicariant barriers operate, the results 489 presented here and elsewhere (Kaefer and Lima 2012; Kaefer et al. 2012; Tsuji-Nishikido et 490 al. 2012) support the notion that the Amazon and Madeira Rivers do play a role in the 491 distribution of the genetic and phenotypic variability of the studied systems. Our conclusions 492 add to a growing body of studies that supported the effectiveness of the Amazon (e.g., 493 Capparella 1988; Hayes and Sewlal 2004), Madeira (e.g., Cohn-Haft 2000; Simões et al. 494 2008; Fernandes et al. 2012) and other Amazonian rivers (e.g., Funk et al. 2007; Ribas et al. 495 2011; Fouquet et al. in press) as vicariant barriers promoting and maintaining evolutionary 496 diversity. In this context, the various studies conducted at the Juruá River, another southern 69 497 tributary of the Amazon, which rejected the river barrier hypothesis (e.g., Patton et al. 1994; 498 Gascon et al. 1996; 1998; 2000; Lougheed et al. 1999) seem to reflect the ineffectiveness of 499 this particular meandering river as a barrier to gene flow rather than reject the river barrier 500 hypothesis for the entire biome. 501 Stochastic and selective forces in evolutionary differentiation 502 In addition to the genetic discontinuities associated with the transposition of the large 503 Amazonian rivers, results from both study systems suggested that a large part of their genetic 504 variation is correlated with geographic distance. This was more evident in A. paleovarzensis, 505 for which Mantel tests indicated a pronounced correlation between genetic and geographic 506 distances together with a non-significant role of the river barrier in genetic differentiation. 507 Distance effects have been reported from phylogeographic studies involving Amazonian frogs 508 (Funk et al. 2007; Amézquita et al. 2009; Simões 2010) and suggest an important role for 509 isolation by distance and genetic drift on their evolutionary diversification. Divergent sexual 510 selection was observed to occur in Amazonian frog species where genetic divergence and 511 reduced gene flow were associated with differential female preferences (Boul et al. 2007; 512 Guerra and Ron 2008). Therefore, a combination of stochastic and selective forces seems to 513 have contributed to the richness of frog lineages observed in Amazonia. This assumption is in 514 line with the idea that neutral and selective processes are non-exclusive explanations for the 515 evolution of a given character (Hansen and Orzack 2005). 516 The acoustic and morphological traits analyzed in this study did not respond to linear 517 geographic distance among sampling localities, as observed with molecular characters. In 518 addition, the discriminant function analyses revealed that a large fraction of the individuals 519 could not be correctly classified in their respective genetic clusters according to their 520 phenotypes. This misclassification may be attributed to a single marker-effect, given that 521 mitochondrial genealogies may exhibit patterns of variation different that those obtained from 522 the analysis of nuclear markers (e.g., Sequeira et al. 2011; Turmelle et al. 2011, but see Creer 523 et al. 2004; Zink and Barrowclough 2008). Nonetheless, the significant correlations between 524 our mitochondrial data and the main geographic barriers of the study areas (river and 525 distance) suggest that the 16S marker was sensitive to the recent history of connectedness 526 among these populations. Therefore, it is reasonable to explore other explanations to the 527 relationship between the evolution of genetic and phenotypic characters. 70 528 The results of this study, together with the observation that there is low among- 529 population variability in phenotypic characters (Kaefer and Lima 2012; Kaefer et al. 2012), 530 indicate a prevailing role of stabilizing selective forces in phenotypic differentiation. In this 531 case, evolutionary constraints may be preventing segregation in morphology and calls among 532 populations. Additionally, the time elapsed since the allopatric condition may not have been 533 sufficient for the accumulation of significant differences. 534 Supporting the first explanation is the observation that well diverged lineages of frogs 535 generally exhibit low phenotypic differentiation (Cherry et al. 1978; Wiens 2008), and that 536 the functional importance of these characters imposes evolutionary constraints (Lougheed et 537 al. 2006). According to the phylogenetic niche conservatism hypothesis, an important and 538 often neglected component of allopatric speciation is the tendency of lineages to maintain 539 their ancestral ecological niche (Wiens 2004). Our results appear to support the occurrence of 540 this phenomenon, as it predicts that closely related lineages are more ecologically similar that 541 would be expected based on their phylogenetic relationships (Losos 2008). This evolutionary 542 conservatism, in which diversification can occur with little phenotypic change (Adams et al. 543 2009) might be related, for example, to the historical failure of researchers in providing 544 reasonable estimates of species richness for many anuran groups based on the classic 545 morphological and biological species concepts (Camargo et al. 2006; Fouquet et al. 2007; 546 Angulo and Reichle 2008; Funk et al. 2012). In both studied systems, bioacoustic characters 547 did not echo fine-scale phylogeographic structure, supporting the notion that mate-recognition 548 signals are subject to severe constraining or diversifying pressures and therefore are not 549 closely related to geographic distance (Gerhardt and Huber 2002). In fact, behavioral traits are 550 expected to be relatively evolutionarily malleable (Blomberg et al. 2003). However, 551 exceptions to this pattern were reported (Ryan et al. 1996; Wycherley et al. 2002; Pröhl et al. 552 2007; Amézquita et al. 2009), depicting the large array of evolutionary forces underlying 553 geographic divergence, as well as the idiosyncratic nature of their prevalence. 554 In favour of the second interpretation is the observation that the only the more 555 differentiated system, composed by A. nidicola and A. masniger, exhibited significant 556 differences in summarized acoustic parameters between riversides. It suggests that the 557 diversifying effects of genetic drift assumed prominence in ancient, isolated lineages simply 558 because of the accumulation of mutations over a longer time span. This interpretation also 559 assumes that, under non-selective evolutionary forces, neutrally evolving genotypic characters 560 diverge first, and that more time is needed for phenotypic differences to arise. 71 561 Conclusions 562 This study indicated that the distribution of genetic variability in three species of 563 neotropical frogs is determined by a combination of the effects of isolation-by-distance and 564 the transposition of large Amazonian rivers acting as vicariant barriers. Thus, we confirm our 565 expectation that genetic drift plays a central role in the emergence of evolutionary 566 differentiation that may lead to the formation of new species. In addition, the current 567 phenotypic divergence was loosely related with genetic differentiation and did not 568 successfully predict assignment of individuals to genetic groups. These results argue against 569 the hypothesis that the phenotypic differentiation is related to the genetic groupings that 570 compose the study systems, and are surprising when considering the high levels of genetic 571 structuring observed. Together with recent studies conducted on these model organisms 572 (Kaefer and Lima 2012; Kaefer et al. 2012; Tsuji-Nishikido et al. 2012), we suggest a 573 prevailing role of stabilizing selective forces in phenotypic differentiation. In this case, 574 evolutionary constraints, which have no effect on neutrally evolving molecular markers, may 575 be preventing phenotypic differentiation. 576 577 Ethical standards 578 This study complies with the current Brazilian laws and was allowed by RAN- 579 ICMBio/IBAMA (licences 13777-2, 18516-1, 20065-2, and 21950-1). 580 581 Conflict of interest 582 The authors declare that they have no conflict of interest. 583 584 Acknowledgements 585 We thank Anelise Montanarin, Francisco C. de Freitas, Irene da S. Melo, Maria A. 586 Carvalho, Moisés da S. Melo and Raimundo N. Amorim for fieldwork assistance; Daniela 587 Leroy e Waleska Gravena for help in laboratory procedures; Adolfo Amézquita, Adrian 72 588 Garda, Andrew J. Crawford, Janet W. Reid, Jeffrey Podos, Luciana K. Erdtmann, Marcelo 589 Menin, Marina Anciães, Pedro Ivo Simões and Tomas Hrbek for valuable suggestions during 590 this study. We also thank the Brazilian Conselho de Desenvolvimento Científico e 591 Tecnológico for financial support (CNPq-CT Amazônia 553997/2006-8 and 575572/2008-6). 592 593 References 594 Adams, D. C., Berns, C. M., Kozak, K. H. & Wiens, J. (2009). Are rates of species 595 diversification correlated with rates of morphological evolution? Proceedings of the Royal 596 Society B, 276, 2729–2738. 597 Amézquita, A., Lima, A. P., Jehle, R., Castellanos, L., Ramos, O., Crawford, A. J., Gasser, H. 598 & Hödl, W. (2009). Calls, colours, shapes, and genes: A multi–trait approach to the study of 599 geographic variation in the Amazonian frog Allobates femoralis. Biological Journal of the 600 Linnean Society, 98, 826–838. 601 Anderson, M. J. & Legendre, P. (1999). An empirical comparison of permutation methods for 602 tests of partial regression coefficients in a linear model. Journal of Statistical Computation 603 and Simulation, 62, 271–303. 604 Angulo, A. & Reichle, S. (2008). Acoustic signals, species diagnosis, and species concepts: 605 the case of a new cryptic species of Leptodactylus (Amphibia, Anura, Leptodactylidae) from 606 the Chapare region, Bolivia. Zoological Journal of the Linnean Society, 152, 58–77. 607 Antonelli, A., Quijada-Mascareñas, A., Crawford, A. J., Bates, J. M., Velazco, P. M & Wüster 608 W. (2010). Molecular studies and phylogeography of Amazonian tetrapods and their relation 609 to geological and climatic models. In C Hoorn, F. P. Wesselingh (Eds.), Amazonia, landscape 610 and species evolution (pp. 386–403). Oxford: Blackwell Publishing. 611 Avise, J. C. (2004). Molecular Markers, Natural History and Evolution. 2nd edition. 612 Sunderland: Sinauer Associates Inc. 613 Avise, J. C. (2000). Phylogeography: the history and formation of species. Cambridge: 614 Harvard University Press. 73 615 Ayres, J. M. & Clutton-Brock, T. H. (1992). River boundaries and species range size in 616 Amazonian primates. The American Naturalist, 140, 531–537. 617 Blomberg, S. P., Garland Jr. T. & Ives, A. R. (2003). Testing for phylogenetic signal in 618 comparative data: behavioral traits are more labile. Evolution, 57, 717–745. 619 Bonnet, E. & Van de Peer, Y. (2002). ZT: a software tool for simple and partial Mantel tests. 620 Journal of Statistical Software, 7, 1–12. 621 Boul, K. E., Funk, W. C., Darst, C. R., Cannatella, D. C. & Ryan, M. J. (2007). Sexual 622 selection drives speciation in an Amazonian frog. Proceedings of the Royal Society B, 274, 623 399–406. 624 Caldwell, J. P. & Lima, A. P. (2003). A new Amazonian species of Colostethus (Anura: 625 Dendrobatidae) with a nidicolous tadpole. Herpetologica, 59, 219–234. 626 Camargo, A., de Sá, R. & Heyer, W. R. (2006). Phylogenetic analyses of mtDNA sequences 627 reveal three cryptic lineages in the widespread neotropical frog Leptodactylus fuscus 628 (Schneider, 1799) (Anura, Leptodactylidae). Biological Journal of the Linnean Society, 87, 629 325–341. 630 Campbell, P., Pasch, B., Pino, J. L., Crino, O. L., Phillips, M. & Phelps, S. M. (2010). 631 Geographic variation in the songs of neotropical singing mice: testing the relative importance 632 of drift and local adaptation. Evolution, 64, 1955–1972. 633 Capparella, A. P. (1988). Genetic variation in Neotropical birds: implications for the 634 speciation process. In H. Ouellet (Ed.), Acta XIX Congressus Internationalis Ornithologici 635 (pp. 1658–1664). Ottawa: University of Ottawa Press. 636 Charif, R. A., Clark, C. W. & Fristrup, K. M. (2004). Raven 1.2 User’s Manual. Ithaca: 637 Cornell Laboratory of Ornithology. 638 Charif R. A., Waack, A. M. & Strickman, L. M. (2008). Raven Pro 1.3 User’s Manual. 639 Ithaca: Cornell Laboratory of Ornithology. 640 Cherry, L. M., Case, S. M. & Wilson, A. C. (1978). Frog perspective on the morphological 641 difference between humans and chimpanzees. Science, 200, 209–211. 74 642 Clement, M., Posada, D. & Crandall, K. A. (2000). TCS: a computer program to estimate 643 gene genealogies. Molecular Ecology, 9, 1657–1659. 644 Cohn-Haft, M. (2000). A case study in amazonian biogeography: Vocal and DNA sequence 645 variation in Hemitriccus flycatchers. PhD thesis. Luisiana State University. 646 Colwell, R.K. (2000). A barrier runs through it… or maybe just a river. Proceedings of the 647 National Academy of Sciences, 97, 13470–13472. 648 Corander, J., Marttinen, P., Sirén, J. & Tang, J. (2008). Enhanced Bayesian modelling in 649 BAPS software for learning genetic structures of populations. BMC Bioinformatics, 9, 539. 650 Costa, J. B. S., Bemerguy, R. L., Hasui, Y. & Borges, M. S. (2001). Tectonics and 651 paleogeography along the Amazon River. Journal of South American Earth Sciences, 14, 652 335–347. 653 Coyne, J. A. & Orr, H. A. (2004). Speciation. Sunderland: Sinauer Associates Inc. 654 Crawford, A. J. (2003). Huge populations and old species of Costa Rican and Panamanian dirt 655 frogs inferred from mitochondrial and nuclear gene sequences. Molecular Ecology, 12, 2525– 656 2540. 657 Crawford, A. J., Lips, K. R. & Bermingham, E. (2010). Epidemic disease decimates 658 amphibian abundance, species diversity, and evolutionary history in the highlands of central 659 Panama. Proceedings of the National Academy of Sciences, 107, 13777–13782. 660 Creer, S., Thorpe, R. S., Malhotra, A., Chou, W. H. & Stenson, A. G. (2004). The utility of 661 AFLPs for supporting mitochondrial DNA phylogeographical analyses in the Taiwanese 662 bamboo viper, Trimeresurus stejnegeri. Journal of Evolutionary Biology, 17, 100–107. 663 Crisci, J. V., Katinas, L., & Posadas, P. (2003). Historical biogeography: An Introduction. 664 Cambridge: Harvard University Press. 665 Dall, S. R. X. (1997). Behaviour and speciation. Trends in Ecology and Evolution , 12, 209– 666 210. 667 Dayrat, B. (2005). Towards integrative taxonomy. Biological Journal of the Linnean Society, 668 85, 407–415. 75 669 de Queiroz, K. (2007). Species concepts and species delineation. Systematic Biology, 56, 879– 670 886. 671 Diniz-Filho, J. A. F., Telles, M. P. C., Bonatto, S. L., Eizirik, E., Freitas, T. R. O., de Marco, 672 P., Santos, F. R., Solé–Cava, A. M. & Soares, T. N. (2008). Mapping the evolutionary 673 twilight zone: molecular markers, populations and geography. Journal of Biogeography, 35, 674 753–763. 675 Drummond, A. J. & Rambaut, A. (2007). BEAST: Bayesian evolutionary analysis by 676 sampling trees. BMC Evolutionary Biology, 7, 214. 677 Elmer K. R., Dávila, J.A. & Lougheed, S.C. (2007). Cryptic diversity and deep divergence in 678 an upper Amazonian leaflitter frog, Eleutherodactylus ockendeni. BMC Evolutionary Biology, 679 7, 247. 680 Erdtmann, L. & Amézquita, A. (2009). Differential evolution of advertisement call traits in 681 Dart–Poison Frogs (Anura: Dendrobatidae). Ethology, 115, 801–811. 682 Excoffier, L. & Lischer, H. E. L. (2010). Arlequin suite ver 3.5: A new series of programs to 683 perform population genetics analyses under Linux and Windows. Molecular Ecology 684 Research, 10, 564–567. 685 Excoffier, L., Smouse, P. & Quattro, J. (1992). Analysis of molecular variance inferred from 686 metric distances among DNA haplotypes: application to human mitochondrial DNA 687 restriction data. Genetics, 131, 479–491. 688 Fernandes, A. M., Wink, M. & Aleixo, A. (2012). Phylogeography of the chestnut tailed 689 antbird (Myrmeciza hemimelaena) clarifies the role of rivers in Amazonian biogeography. 690 Journal of Biogeography, doi:10.1111/j.1365-2699.2012.02712.x 691 Figueiredo, J., Hoorn, C., van der Ven, P. & Soares, E. (2009). Late Miocene onset of the 692 Amazon River and the Amazon deep-sea fan: Evidence from the Foz do Amazonas Basin. 693 Geology, 37, 619–622. 694 Fouquet, A., Gilles, A., Vences, M., Marty, C., Blanc, M. & Gemmell, N. J. (2007). 695 Underestimation of species richness in Neotropical frogs revealed by mtDNA analyses. PLoS 696 One, 10, e1109. 76 697 Fouquet, A., Noonan, B. P., Rodrigues, M. T., Pech, N., Gilles, A. & Gemmell, N. J.: 698 Multiple Quaternary Refugia in the Eastern Guiana Shield Revealed by Comparative 699 Phylogeography of 12 Frog Species. Systematic Biology, in press. 700 Fouquet, A., Recoder, R., Teixeira–Jr, M., Cassimiro, J., Amaro, R. C., Camacho, A., 701 Damasceno, R., Carnaval, A. C., Moritz, C. & Rodrigues, M. T. (2012). Molecular phylogeny 702 and morphometric analyses reveal deep divergence between Amazonia and Atlantic Forest 703 species of Dendrophryniscus. Molecular Phylogenetics and Evolution, 62, 826–838. 704 Fu, Y. X. (1997). Statistical tests of neutrality of mutations against population growth, 705 hitchhiking and background selection. Genetics, 147, 915–925. 706 Funk, D. J. & Omland, K. E. (2003). Species-level paraphyly and polyphyly: frequency, 707 causes, and consequences, with insights from animal mitochondrial DNA. Annual Review of 708 Ecology, Evolution, and Systematics, 34, 397–423. 709 Funk, W. C., Caldwell, J. P., Peden, C. E., Padial, J. M., de la Riva, I. & Cannatella, D. C. 710 (2007). Tests of biogeographic hypotheses for diversification in the Amazonian forest frog, 711 Physalaemus petersi. Molecular Phylogenetics and Evolution, 44, 825–837. 712 Funk, W. C., Caminer, M. & Ron, S. R. (2012). High levels of cryptic species diversity 713 uncovered in Amazonian frogs. Proceedings of the Royal Society B, 279, 1806–1814. 714 Gascon, C., Lougheed, S. C. & Bogart, J. P. (1996). Genetic and morphological variation in 715 Vanzolinius discodactylus: a direct test of the riverine barrier hypothesis. Biotropica, 28, 376– 716 387. 717 Gascon, C., Lougheed, S. C. & Bogart, J. P. (1998). Patterns of genetic population 718 differentiation in four species of Amazonian frogs: A test of the Riverine Barrier Hypothesis. 719 Biotropica, 30, 104–119. 720 Gascon, C., Malcolm, J. R., Patton, J.L., Silva, M. N. F., Bogart, J.P., Lougheed, S.C., Peres, 721 C.A., Neckel, S. & Boag P.T. (2000). Riverine barriers and the geographic distribution of 722 Amazonian species. Proceedings of the National Academy of Sciences, 97, 13672–13677. 723 Gerhardt. H.C. & Huber, F. (2002). Acoustic Communication in Insects and Anurans: 724 Common Problems and Diverse Solutions. Chicago: University of Chicago Press. 77 725 Goicochea, N., De La Riva, I. & Padial, J. M. (2009). Recovering phylogenetic signal from 726 frog mating calls. Zoologica Scripta, 39, 141–154. 727 Grant, T., Frost, D. R., Caldwell, J. P., Gagliardo, R., Haddad, C. F. B., Kok, P. J. R., Means, 728 B. D, Noonan, B. P., Schargel, W. & Wheeler, W. C. (2006). Phylogenetic systematics of 729 dart–poison frogs and their relatives (Anura: Athesphatanura: Dendrobatidae). Bulletin of the 730 American Museum of Natural History, 299, 1–262. 731 Groeneveld, L. F., Weisrock, D. W., Rasoloarison, R. M., Yoder, A. D. & Kappeler, P. M. 732 (2009). Species delimitation in lemurs: multiple genetic loci reveal low levels of species 733 diversity in the genus Cheirogaleus. BMC Evolutionary Biology, 9, 30. 734 Guerra, M. A. & Ron, S. R. (2008). Mate choice and courtship signal differentiation promotes 735 speciation in an Amazonian frog. Behavioral Ecology, 19, 1128–1135. 736 Guillot, G., Renaud, S., Ledevin, R., Michaux, J. & Claude, J. (2012). A unifying model for 737 the analysis of phenotypic, genetic and geographic data. Systematic Biology, doi: 738 10.1093/sysbio/sys038. 739 Haffer, J. (1969). Speciation in Amazonian Forest Birds. Science, 165, 131–137. 740 Hall, T. A. (1999). BioEdit: a user-friendly biological sequence alignment editor and analysis 741 program for Windows 95/98/NT. Nucleic Acids Symposium Series, 41, 95–98. 742 Hansen, T. F. & Orzack, S.H. (2005). Assessing current adaptation and phylogenetic inertia as 743 explanations of trait evolution: the need for controlled comparisons. Evolution, 59, 2063– 744 2072. 745 Harpending, R. C. (1994). Signature of ancient population growth in a low–resolution 746 mitochondrial DNA mismatch distribution. Human Biology, 66, 591–600. 747 Hayes, F.E. & Sewlal, J.N. (2004). The Amazon River as a dispersal barrier to passerine 748 birds: effects of river width, habitat and taxonomy. Journal of Biogeography, 31, 1809–1818. 749 Hoorn, C. (1994). An environmental reconstruction of the palaeo-Amazon River system 750 (Middle-Late Miocene, NW 751 Palaeoecology, 112, 187–238. Amazonia). Palaeogeography, Palaeoclimatology, 78 752 Hoorn, C., Wesselingh, F.P., ter Steege, H., Bermudez, M. A., Mora, A., Sevink, J., 753 Sanmartín, I., Sanchez-Meseguer, A., Anderson, C. L., Figueiredo, J.P., Jaramillo, C., Riff, 754 D., Negri, F. R., Hooghiemstra, H., Lundberg, J., Stadler, T., Särkinen, T. & Antonelli, A. 755 (2010). Amazonia through time: Andean uplift, climate change, landscape evolution, and 756 biodiversity. Science, 330, 927–931. 757 Hubert, N., Meyer, C. P., Bruggemann, H. J., Guérin, F., Komeno, R. J. L., Espiau, B., 758 Causse, R., Williams, J. T. & Planes, S. (2012). Cryptic diversity in Indo–Pacific coral–reef 759 fishes revealed by DNA–barcoding provides new support to the centre–of–overlap 760 hypothesis. PloS One, 7, e28987. 761 Hutchison, D. W. & Templeton, A. R. (1999). Correlation of pairwise genetic and geographic 762 distance measures: inferring the relative influences of gene flow and drift on the distribution 763 of genetic variability. Evolution, 53, 1898–1914. 764 Irion, G. & Kalliola, R. (2010). Long–term landscape development processes in Amazonia. In 765 C Hoorn, F. P. Wesselingh (Eds.), Amazonia, landscape and species evolution (pp. 185–197). 766 Oxford: Blackwell Publishing. 767 Kaefer, I. L. & Lima, A. P. (2012). Sexual signals of the Amazonian frog Allobates 768 paleovarzensis: geographic variation and stereotypy of acoustic traits. Behaviour, 149, 15–33. 769 Kaefer, I. L., Montanarin, A., Costa, R. S. & Lima, A. P. Temporal patterns of reproductive 770 activity and site attachment of the Brilliant–thighed Frog Allobates femoralis from Central 771 Amazonia. Journal of Herpetology, in press. 772 Kaefer, I.L., Tsuji-Nishikido, B. M. & Lima, A. P. (2012). Beyond the river: underlying 773 determinants of population acoustic signal variability in Amazonian direct–developing 774 Allobates (Anura: Dendrobatoidea). Acta Ethologica, doi:10.1007/s10211-012-0126-0. 775 Kimura, M. (1980). A simple method for estimating evolutionary rate of base substitutions 776 through comparative studies of nucleotide sequences. Journal of Molecular Evolution, 16, 777 111–120. 778 Lemmon, E. M. (2009). Diversification of conspecific signals in sympatry: geographic 779 overlap drives multidimensional reproductive character displacement in frogs. Evolution, 63, 780 1155–1170. 79 781 Librado, P. & Rozas, J. (2009). DnaSP v5: A software for comprehensive analysis of DNA 782 polymorphism data. Bioinformatics, 25, 1451–1452. 783 Lima, A.P., Caldwell, J. P., Biavati, G. & Montanarin, A. (2010). A new species of Allobates 784 (Anura: Aromobatidae) from paleovárzea forest in Amazonas, Brazil. Zootaxa, 2337, 1–17. 785 Losos, J. B. (2008). Phylogenetic niche conservatism, phylogenetic signal and the relationship 786 between phylogenetic relatedness and ecological similarity among species. Ecology Letters, 787 11, 995–1003. 788 Lougheed, S. C., Austin, J. D., Bogart, J. P., Boag, P.T. & Chek, A. A. (2006). Multi– 789 character perspectives on the evolution of intraspecific differentiation in a neotropical hylid 790 frog. BMC Evolutionary Biology, 6, 23. 791 Lougheed, S. C., Gascon, C., Jones, D. A., Bogart, J. P. & Boag, P. T. (1999). Ridges and 792 rivers: a test of competing hypothesis of Amazonian diversification using a dart–poison frog 793 (Epipedobates femoralis). Proceedings of the Royal Society B, 266, 1829–1835. 794 Lynch, M. & Crease, T. J. (1990). The analysis of population survey data on DNA sequence 795 variation. Molecular Biology and Evolution, 7, 377–394. 796 Mantel, N. (1967). The detection of disease clustering and a generalized regression approach. 797 Cancer Research, 27, 209–220. 798 McKay, B. D. & Zink, R. M. (2009). The causes of mitochondrial DNA gene tree paraphyly 799 in birds. Molecuar Phylogenetics and Evolution, 54, 647–650. 800 Morales, V. R. (2002). Sistematica y biogeografía del grupo trilineatus (Amphibia, Anura, 801 Dendrobatidae, Colostethus), con descripción de once nuevas especies. Publicación de la 802 Asociación Amigos Doñana, 13, 1–59. 803 Moritz, C., Patton, J. L., Schneider, C. J. & Smith, T. B. (2000). Diversification of rainforest 804 faunas: An integrated molecular approach. Annual Review of Ecology and Systematics, 31, 805 533–563. 806 Nosil, P., Harmon, L. J. & Seehausen, O. (2009). Ecological explanations for (incomplete) 807 speciation. Trends in Ecology and Evolution, 24, 145–156. 80 808 Palumbi, S. R. (1996). Nucleic acids II: the polymerase chain reaction. In D. M. Hillis, C. 809 Moritz, B. K. Mable (Eds.), Molecular Systematics (pp. 205–247). Sunderland: Sinauer 810 Associates Inc. 811 Panhuis, T. M., Butlin, R., Zuk, M. & Tregenza, T. (2001). Sexual selection and speciation. 812 Trends in Ecology and Evolution, 16, 364–371. 813 Patton, J. L., Silva, M. N. F. & Malcolm, J. R. (1994). Gene genealogy and differentiation 814 among arboreal spiny rats (Rodentia: Echymidae) of the Amazon Basin: a test of the riverine 815 barrier hypothesis. Evolution, 48, 1314–1323. 816 Pröhl, H. (2005). Territorial behavior in dendrobatid frogs. Journal of Herpetology, 39, 354– 817 365. 818 Pröhl, H., Hagemann, S., Karsch, J. & Höbel, G. (2007). Geographic variation in male sexual 819 signals in Strawberry Poison Frogs (Dendrobates pumilio). Ethology, 113, 825–837. 820 Pröhl, H., Koshy, R. A., Mueller, U. G., Rand, A. S. & Ryan, M. J. (2006). Geographic 821 variation of genetic and behavioral traits in northern and southern túngara frogs. Evolution, 822 60, 1669–1679. 823 Rambaut, A. & Drummond, A. J. (2007). Tracer v. 1.5. [http://beast.bio.ed.ac.uk/Tracer]. 824 Accessed 04 May 2012. 825 Ramos–Onsins, S. E. & Rozas, J. (2002). Statistical properties of new neutrality tests against 826 population growth. Molecular Biology and Evolution, 19, 2092–2100. 827 Ribas, C. C., Aleixo, A., Nogueira, A. C. R., Miyaki, C. Y. & Cracraft, J. (2011). A 828 palaeobiogeographic model for biotic diversification within Amazonia over the past three 829 million years. Proceedings of the Royal Society B, 279, 681–689. 830 Rosen, D. E. (1978). Vicariant patterns and historical explanation in biogeography. Systematic 831 Zoology, 27, 159–188. 832 Rossetti, D. F., de Toledo, P. M. & Góes, A. M. (2005). New geological framework for 833 Western Amazonia (Brazil) and implications for biogeography and evolution. Quaternary 834 Research, 63, 78–89. 81 835 Ryan, M. J., Rand, A. S. & Weigt, L. A. (1996). Allozyme and advertisement call variation in 836 the túngara frog, Physalaemus pustulosus. Evolution, 50, 2435–2453. 837 Santini, F., Miglietta, M. P. & Faucci, A. (2012). Where are we now? An introduction to a 838 special issue on speciation. Evolutionary Biology, 39: 141-147. 839 Santos, J. C., Coloma, L. A., Summers, K., Caldwell, J. P., Ree, R. & Cannatella, D. C.: 840 (2009). Amazonian amphibian diversity is primarily derived from Late Miocene Andean 841 lineages. PLoS Biology, 7, e1000056. 842 Santos, S., Hrbek, T., Farias, I. P., Schneider, H. & Sampaio, I. (2006). Population genetic 843 structuring of the king weakfish, Macrodon ancylodon (Sciaenidae), in Atlantic coastal waters 844 of South America: deep genetic divergence without morphological change. Molecular 845 Ecology, 15, 4361–4373. 846 Schneider, S. & Excoffier, L. (1999). Estimation of demographic parameters from the 847 distribution of pairwise differences when the mutation rates vary among sites: Application to 848 human mitochondrial DNA. Genetics, 152, 1079–1089. 849 Sequeira, F., Sodré, D., Ferrand, N., Bernardi, J., Sampaio, I., Schneider, H. & Vallinoto, M. 850 (2011). Hybridization and massive mtDNA unidirectional introgression between the closely 851 related Neotropical toads Rhinella marina and R. schneideri inferred from mtDNA and 852 nuclear markers. BMC Evolutionary Biology, 11, 264. 853 Silva, M. N. F & Patton, J. L. (1998). Molecular phylogeography and the evolution and 854 conservation of mammals. Molecular Ecology, 7, 475–486. 855 Simões, P. I. (2010). Diversificação do complexo Allobates femoralis (Anura, Dendrobatidae) 856 em florestas da Amazônia brasileira: desvendando padrões atuais e históricos. PhD thesis. 857 Instituto Nacional de Pesquisas da Amazônia, Manaus. 858 Simões, P. I., Lima, A. P., Magnusson, W. E., Hödl, W. & Amézquita, A. (2008). Acoustic 859 and morphological differentiation in the frog Allobates femoralis: Relationships with the 860 Upper Madeira River and other potential geological barriers. Biotropica, 40, 607–614. 861 Sioli, H. (1984). The Amazon: Limnology and Landscape Ecology of a Mighty Tropical River 862 and its Basin. Dordrecht: Dr W. Junk Publisher. 82 863 Smouse, P. E., Long, J. C. & Sokal, R. R. (1986). Multiple regression and correlation 864 extensions of the Mantel test of matrix correspondence. Systematic Zoology, 35, 727–732. 865 Solomon, S. E., Bacci, M., Martins, J., Vinha, G. G. & Mueller, U. G. (2008). 866 Paleodistributions and comparative molecular phylogeography of leafcutter ants (Atta spp.) 867 provide new insight into the origins of Amazonian diversity. PLoS One, 3, e2738. 868 Tajima, F. (1989). Statistical method for testing the neutral mutation hypothesis by DNA 869 polymorphism. Genetics, 123, 585–595. 870 Tamura, K., Peterson, D., Peterson, N., Stecher, G., Nei, M. & Kumar, S. (2011). MEGA5: 871 Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, 872 and maximum parsimony methods. Molecular Biology and Evolution, 28, 2731–2739. 873 Telles, M. P. C., Silva, R. S. M., Chaves, L. J., Coelho, A. S. G & Diniz–Filho, J. A. F. 874 (2001). Divergência entre subpopulações de cagaiteira (Eugenia dysenterica) em resposta a 875 padrões edáficos e distribuição espacial. Pesquisa Agropecuária Brasileira, 36, 1387–1394. 876 Templeton, A. R., Crandall, K. A. & Sing, C. F. (1992). A cladistic analysis of phenotypic 877 associations with haplotypes inferred from restriction endonuclease mapping and DNA 878 sequence data. III. Cladogram estimation. Genetics, 132, 619–633. 879 Thompson, J. D., Higgins, D. G. & Gibson, T. J. (1994). Improved sensitivity of profile 880 searches through the use of sequence weights and gap excision. Computer Applications in the 881 Biosciences, 10, 19–29. 882 Tsuji-Nishikido, B., Kaefer, I. L., Freitas, F. C., Menin, M. & Lima, A. P. (2012). Significant 883 but not diagnostic: Differentiation through morphology and calls in the Amazonian frogs 884 Allobates nidicola and A. masniger. Herpetological Journal, 22, 105–114. 885 Turmelle, A. S., Kunz, T. H. & Sorenson, M. D. (2011). A tale of two genomes: contrasting 886 patterns of phylogeographic structure in a widely distributed bat. Molecular Ecology, 20, 887 357–375. 888 Vences, M., Kosuch, J., Lötters, S., Widmer, A., Jungfer, K., Köhler, J. & Veith, M. (2000). 889 Phylogeny and classification of Poison Frogs (Amphibia: Dendrobatidae), based on 890 mitochondrial 16S and 12S ribosomal RNA gene sequences. Molecular Phylogenetics and 891 Evolution, 15, 34–40. 83 892 Vences, M., Thomas, M., Meijden, A., Chiari, Y. & Vieites, D. R. (2005). Comparative 893 performance of the 16S rRNA gene in DNA barcoding of amphibians. Frontiers in Zoology, 894 2, 5. 895 Vences, M. & Wake, D. B. (2007). Speciation, species boundaries and phylogeography of 896 amphibians. In HH Heatwole, M. Tyler (Eds.), Amphibian Biology. Volume 6: Systematics 897 (pp. 2613–2669). Chipping Norton: Surrey Beatty & Sons. 898 Verdade, V. K. & Rodrigues, M. T. (2007). Taxonomic review of Allobates (Anura, 899 Aromobatidae) from the Atlantic Forest, Brazil. Journal of Herpetology, 41, 566–580. 900 Vieites, D. R., Wollenberg, K. C., Adreone, F., Köhler, J., Glaw, F. & Vences, M. (2009). 901 Vast underestimation of Madagascar’s biodiversity evidenced by an integrative amphibian 902 inventory. Proceedings of the National Academy of Sciences, 106, 8267–8272. 903 Wallace, A. R. (1852). On the monkeys of the Amazon. Proceedings of the Zoological 904 Society of London, 20, 107–110. 905 West-Eberhard, M. J. (1989). Phenotypic plasticity and the origins of diversity. Annual 906 Review of Ecology and Systematics, 20, 249–278. 907 Wiens, J. J. (2004). Speciation and ecology revisited: phylogenetic niche conservatism and 908 the origin of species. Evolution, 58, 193–197. 909 Wiens, J. J. (2007). Species delimitation: New approaches for discovering diversity. 910 Systematic Biology, 56, 875–878. 911 Wiens, J. J. (2008). Systematics and herpetology in the age of genomics. Bioscience, 58, 297– 912 307. 913 Wiley, E. O. (1978). The evolutionary species concept reconsidered. Systematic Zoology, 27, 914 17–26. 915 Wright, S. (1951). The genetical structure of populations. Annals of Human Genetics, 15, 916 323–354. 917 Wycherley, J., Doran, S. & Beebee, T. J. C. (2002). Male advertisement call characters as 918 phylogeographical indicators in European water frogs. Biological Journal of the Linnean 919 Society, 77, 355–365. 84 920 Zeisset, I. & Beebee, T. J. C. (2008). Amphibian phylogeography: a model for understanding 921 historical aspects of species distributions. Heredity, 101, 109–119. 922 Zhang, D. X. & Hewitt, G. M. (2003). Nuclear DNA analyses in genetic studies of 923 populations: Practice, problems and prospects. Molecular Ecology, 12, 563–584. 924 Zink, R. M. & Barrowclough, G. F. (2008). Mitochondrial DNA under siege in avian 925 phylogeography. Molecular Ecology, 17, 2107–2121. 926 Zink, R. M., Blackwell-Rago, R.C. & Ronquist, F. (2000). The shifting roles of vicariance in 927 biogeography. Proceedings of the Royal Society B, 267, 497–503. 85 Tables Table 1 Study sites in Brazilian Amazonia with respective geographic coordinates and sample sizes. Type localities are indicated by asterisks. Sample sizes are given as the number of Allobates sampled for each class of character. Locality Allobates paleovarzensis 1. Careiro* 2. Janauacá 3. Manaquiri 4. Hiléia 5. Novo Airão 6. Anamã 7. Anori 8. Codajás 9. Unini 10. Barcelos Allobates nidicola 1. Autazes Road Km 12* 2. BR-319, km 260 3. PPBio Manaquiri 4. BR-319 Tupana 5. Vila Gomes Allobates masniger 6. Borba 7. Novo Aripuanã 8. Road to Apuí 9. Jacareacanga 10. PARNA da Amazônia* Total Coordinates Morphology Acoustics mtDNA 3°22'26.30"S 59°52'6.40"W 3°28'45.00"S 60°12'11.00"W 3°25'58.00"S 60°36'43.80"W 3°11'52.80"S 60°26'33.00"W 3° 6'5.50"S 60°47'16.10"W 3°30'52.10"S 61°27'14.00"W 3°39'39.20"S 61°39'46.10"W 3°46'17.50"S 62°1'13.90"W 1°43'9.50"S 61°54'22.70"W 0°54'57.70"S 62°58'14.30"W 15 12 15 15 14 13 12 12 09 13 15 12 15 15 14 13 12 12 02 13 17 14 15 15 17 13 09 12 18 12 3°28'2.70"S 59°49'9.16"W 4°37'0.43"S 61°14'27.70"W 3°40'31.52"S 60°19'44.85"W 4° 6'2.84"S 60°39'25.40"W 4°19'33.70"S 59°41'0.33"W 10 10 10 10 10 10 10 10 10 10 11 12 14 13 14 4°26'22.00"S 59°37'10.10"W 5° 7'11.10"S 60°15'2.87"W 5°33'12.10"S 60° 7'31.30"W 6°14'55.50"S 57°52'28.50"W 4°32'51.70"S 56°18'13.60"W 10 10 10 10 10 230 10 10 10 10 10 223 15 15 12 14 14 276 86 Table 2 Analyses of molecular variance (AMOVA) based on fragments of the mitochondrial 16S rDNA. The relative distribution of genetic variability is presented according to hierarquical levels of genetic structure in each of the Allobates study systems. Source of variation Between riversides Among populations within riversides Within populations paleovarzensis system nidicola-masniger system Percentage of variation 20.91 59.74 60.37 38.23 18.71 2.02 87 Table 3 Pairwise FST fixation indexes (lower left matrix) and average Kimura-2-parameter genetic distances (upper right matrix). Values were calculated between the sampling localities of each of the Allobates study systems in Brazilian Amazonia. Significant FST values are indicated with asterisks. Locality 1 2 3 4 5 6 7 8 9 10 1 0.529* 0.630* 0.801* 0.750* 0.711* 0.647* 0.674* 0.901* 0.858* 2 0.005 0.671* 0.791* 0.746* 0.707* 0.636* 0.670* 0.892* 0.844* 3 0.003 0.005 0.789* 0.669* 0.634* 0.528* 0.528* 0.918* 0.918* Localitiy 1 2 3 4 5 6 7 8 9 10 1 0.399* -0.018 0.604* 0.917* 0.991* 0.992* 0.986* 0.989* 0.983* 2 0.002 0.354* 0.315* 0.821* 0.975* 0.977* 0.967* 0.972* 0.970* 3 0.000 0.002 0.582* 0.907* 0.990* 0.990* 0.985* 0.988* 0.983* paleovarzensis system 4 5 6 0.005 0.006 0.006 0.007 0.007 0.007 0.002 0.003 0.003 0.001 0.001 0.097 0.001 0.007 0.062 0.133* 0.131* 0.076* 0.015 0.057 -0.072 0.944* 0.918* 0.904* 0.910* 0.876* 0.851* nidicola-masniger system 4 5 6 0.004 0.004 0.039 0.004 0.006 0.041 0.004 0.005 0.039 0.008 0.043 0.820* 0.043 0.969* 0.984* 0.972* 0.985* 0.964* 0.960* 0.977* 0.976* 0.967* 0.981* 0.978* 0.964* 0.977* 0.974* 7 0.007 0.009 0.004 0.002 0.003 0.003 0.058* 0.867* 0.794* 8 0.006 0.007 0.003 0.001 0.002 0.002 0.003 0.885* 0.823* 9 0.016 0.017 0.012 0.014 0.015 0.014 0.016 0.014 0.131* 10 0.014 0.016 0.011 0.012 0.013 0.012 0.014 0.012 0.002 - 7 0.037 0.039 0.038 0.041 0.041 0.010 0.974* 0.978* 0.975* 8 0.033 0.035 0.033 0.035 0.033 0.028 0.026 0.965* 0.960* 9 0.042 0.042 0.042 0.043 0.041 0.034 0.032 0.024 0.965* 10 0.050 0.052 0.050 0.052 0.049 0.040 0.038 0.030 0.040 - 88 Table 4 Descriptive statistics of genetic polymorphism parameters and results of neutrality tests performed on each of the genetic clusters of Allobates. Bayesian Analysis of Genetic Structure was used to estimate the composition of the genetic clusters. n = sample size; h = number of haplotypes; S = number of segregating sites; π = average (± one standard deviation) pairwise distance between samples pertaining to the same cluster; D = Tajima’s D; Fs = Fu’s Fs; R2 = Ramos-Onsins & Roza’s R2. Significant values in neutrality tests are indicated with asterisks. Cluster Localities paleovarzensis system 1 2+3+4+5+6+7+8 2 1+2+3 3 9+10 4 6+8 nidicola-masniger system 1 10 2 1+2+3+4 3 5 4 6 5 7 6 8 7 9 n h S π ± 1 SD D D 95% Fs Fs 95% R2 R2 95% 79 31 30 2 10 7 8 2 13 6 8 1 0.00146 ± 0.00026 0.00307 ± 0.00030 0.00169 ± 0.00048 0.00190 ± 0.00095 -1.972 0.209 -1.696 - <0.001* 0.625 0.017* - -5.945 -1.074 -4.483 - 0.002* 0.281 0.001* - 0.038 0.130 0.066 0.500 0.040* 0.590 0.034* 1.000 14 50 14 15 15 12 14 3 7 2 3 3 4 3 4 8 1 2 3 3 2 0.00172 ± 0.00066 0.00210 ± 0.00037 0.00092 ± 0.00016 0.00071 ± 0.00031 0.00075 ± 0.00046 0.00094 ± 0.00039 0.00053 ± 0.00030 -0.876 -1.006 1.212 -1.001 -1.685 -1.629 -1.480 0.239 0.173 0.906 0.186 0.009* 0.015* 0.028* 0.686 -1.569 1.139 -0.917 -0.830 -2.123 -1.475 0.687 0.211 0.824 0.095 0.107 0.007* 0.033* 0.202 0.081 0.247 0.139 0.180 0.144 0.175 0.707 0.253 0.754 0.193 0.462 0.046* 0.300 89 Table 5 Tests of demographic expansion performed according to genetic clusters of Allobates indicated by BAPS. Values were inferred from the sum of squared deviation (SSD) between observed and expected mismatch distributions, and Harpending’s raggedness index (Hri). Tests were not conducted on Cluster 4 of Allobates paleovarzensis due to the small number of samples that compose it. n = number of samples. The nonsignificant p values of both tests indicate that the null hypothesis of population expansion through time cannot be rejected. Cluster Localities paleovarzensis system 1 2+3+4+5+6+7+8 2 1+2+3 3 9+10 4 6+8 nidicola-masniger system 1 10 2 1+2+3+4 3 5 4 6 5 7 6 8 7 9 n Mismatch obs. Mean Mismatch obs. variance SSD p (SSD) Hri p (Hri) 79 31 30 2 0.771 1.742 0.892 - 0.878 1.382 1.138 - 0.00009 0.00222 0.01234 - 0.976 0.630 0.202 - 0.073 0.051 0.129 - 0.843 0.498 0.241 - 14 50 14 15 15 12 14 0.923 1.127 0.495 0.381 0.400 0.667 0.549 1.272 1.396 0.253 0.277 0.550 0.718 0.361 0.03276 0.00657 0.02076 0.00506 0.01003 0.00110 0.01290 0.220 0.238 0.180 0.577 0.327 0.783 0.328 0.205 0.093 0.244 0.192 0.396 0.104 0.155 0.208 0.192 0.183 0.377 0.560 0.868 0.533 90 Table 6 Simple and partial Mantel tests evaluating correlations between geographic, phenotypic and genetic distances of Allobates between the studied localities in Brazilian Amazonia. Simple Mantel tests are presented as “Matrix 1” X “Matrix 2” and partial Mantel tests are presented as “Matrix 1” X “Matrix 2”.”Covariate matrix”. GenD = genetic distance; GeoD = geographic distance; MorD = morphological distance; AcoD = acoustic distance; AcoDwa = Acoustic distance without body size adjustment; River = riverside (binary variable). Significant correlations are indicated with asterisks. Model GenD x GeoD MorD x GeoD AcoD x GeoD AcoDwa x GeoD GenD x River.GeoD MorD x River.GeoD AcoD x River.GeoD AcoDwa x River.GeoD paleovarzensis system r p 0.761 0.006* -0.094 0.402 -0.029 0.653 0.327 0.117 0.063 0.379 0.175 0.152 -0.089 0.457 0.014 0.448 nidicola-masniger system R P 0.572 <0.001* 0.530 0.094 -0.041 0.456 -0.150 0.311 0.778 0.002* -0.025 0.425 0.082 0.212 0.388 0.014* 91 Figures Fig. 1 Geographic location of sampling sites (A) and haplotype network for Allobates paleovarzensis (B). The network was built from 142 16S rDNA sequences. The size and color of each ellipse indicate the frequency and geographic origin of the individuals bearing that haplotype. White dots and transverse bars represent not sampled (missing) intermediate haplotypes. Sample localities are numbered according to Table 1. Fig. 2 Geographic location of sampling sites (A) and haplotype network for Allobates nidicola and A. masniger (B). The network was built from 134 16S rDNA sequences. The size and color of each ellipse indicate the frequency and geographic origin of the individuals bearing that haplotype. White dots and transverse bars represent not sampled (missing) intermediate haplotypes. Sample localities are numbered according to Table 1. Fig. 3 Barplot from the Bayesian Analysis of Genetic Differentiation. (A) 142 individual 16S rDNA sequences of Allobates paleovarzensis from ten sampling localities. (B) 134 individual 16S rDNA sequences of Allobates nidicola and A. masniger from other ten sampling localities. Distinct colors represent each estimated genetic cluster. Individuals are displayed according to sampling localities, which are numbered as on Table 1. Fig. 4 Timetree of the (A) paleovarzensis and (B) nidicola-masniger systems based on fragments of the 16S rDNA gene. Only unique haplotypes were included. Age estimates and confidence intervals (horizontal bars) of the divergences were obtained via BEAST (Bayesian Evolutionary Analysis Sampling Trees). Posterior probabilities of the older nodes are presented. Terminals were colored according to their localities of origin, which are depicted in Fig. 1 and 2. Black and white terminals are outgroups. The black and grey vertical bars indicate the origin of each haplotype according to the riverside. 92 Fig. 1 93 Fig. 2 94 Fig. 3 95 Fig. 4 96 Electronic Supplementary Material Online Resource 1 Advertisement call measurements of individuals of Allobates paleovarzensis in each study locality in Brazilian Amazonia. Values are presented as mean (above) and standard deviation (below). Call traits are: Note rate (NR, in notes/s); Note duration (ND, in s); Internote interval, as the silent interval between two consecutive notes of a call (InI, in s); Call rate (CR, in calls/s); Call duration (CD, in s); Intercall interval, as the silent interval between two consecutive calls (IcI, in s); Maximum (peak) frequency, as the frequency of higher intensity calculated for an entire note by a power spectrum function of Raven Pro 1.3 (MF, in Hz); Lowest frequency (LF, in Hz); Highest frequency (HF, in Hz); Note modulation, as the difference between the highest and the lowest frequencies of the call (NM, in Hz). Sampling locality codes, sample sizes and respective geographic coordinates are given on Table 1. Online Resource 2 Advertisement call measurements of individuals of Allobates nidicola and A. masniger in each study locality in Brazilian Amazonia. Values are presented as mean (above) and standard deviation (below). Call traits are: Note duration (ND, in s); Interval between notes (IN, in s); Lowest frequency (LF, in Hz); Highest frequency (HF, in Hz); Note modulation, calculated as the difference between the maximum and minimum frequencies (NM, in Hz); and Peak frequency (PF, in Hz). Sampling locality codes, sample sizes and respective geographic coordinates are given on Table 1. Online Resource 3 Morphometric measurements (in mm) of individuals of Allobates paleovarzensis in each study locality in Brazilian Amazonia. Values are presented as mean (above) and standard deviation (below). Morphometric traits are: Snout-vent length (SVL); head length, from angle of jaws to tip of snout (HL); Head width at level of angle of jaws (HW); Snout length, from anterior corner of eye to tip of snout (SL); Eye to nostril distance, from anterior corner of eye to centre of nostril (EN); Inter-nostril distance (IN); Eye length, from anterior to posterior corner (EL); Inter-orbital distance (IO); Diameter of tympanum (TYM); Forearm length, from proximal edge of palmar tubercle to outer edge of flexed elbow (FAL); Upper arm length, from trunk insertion to outer edge of flexed elbow (UAL); Hand length, from proximal edge of palmar tubercle to tip of fingers I, II, III and IV (HAND1, HAND2, HAND3 and HAND4); Width of disc on finger III (WFD); Tibia length, from outer edge of flexed knee to heel (TL); Foot length, from proximal edge of outer metatarsal tubercle to tip of toe IV (FL); Femur length (LL); Diameter of palmar tubercle (DPT); Width of tenar 97 tubercle (WTT); Width of disc on toe IV (WTD); Width of finger III (WPF). Sampling locality codes, sample sizes and respective geographic coordinates are given on Table 1. Online Resource 4 Morphometric measurements (in mm) of individuals of Allobates nidicola and A. masniger in each study locality in Brazilian Amazonia. Values are presented as mean (above) and standard deviation (below). Morphometric traits are described in Online Resource S3. Sampling locality codes, sample sizes and respective geographic coordinates are given on Table 1. Online Resource 5 Distribution of 16S rDNA haplotypes of Allobates among 20 sampled localities in Brazilian Amazonia. Collection numbers of vouchers (INPA-H) and GenBank accession numbers are provided. Sampled localities are numbered according to Table 1. Online Resource 6 Mismatch distributions obtained from pairwise nucleotide site differences among 16S rDNA sequences composing the clusters of the (A) paleovarzensis and the (B) nidicola-masniger systems. Clusters were determined via Bayesian Analysis of Genetic Differentiation. Pairwise differences were not calculated for Cluster 4 of Allobates paleovarzensis due to the small number of samples that compose it. Online Resource 7 Classification matrix based on the discriminant function analysis, using phenotypic data from males of Allobates paleovazensis pertaining to four genetic clusters. Clusters are coded from C1 to C4, according to Table 4. Body-shape traits were used to discriminate individuals based on morphology. Acoustic measurements were adjusted for temperature and for both temperature and body size through linear regressions. The number and percentage of individuals correctly assigned to each genetic cluster are indicated. Online Resource 8 Classification matrix based on the discriminant function analysis, using phenotypic data from males of Allobates nidicola and A. masniger pertaining to seven genetic clusters. Clusters are coded from C1 to C7, according to Table 4. Body-shape traits were used to discriminate individuals based on morphology. Acoustic measurements were adjusted for temperature and for both temperature and body size through linear regressions. The number and percentage of individuals correctly assigned to each genetic cluster are indicated. 98 Online Resource 1 Locality 1 2 3 4 5 6 7 8 9 10 NR ND 6.24 ±0.99 5.47 ±0.85 5.45 ±0.90 4.60 ±0.83 5.21 ±0.87 6.57 ±1.20 6.97 ±1.30 7.15 ±1.74 6.53 ±0.74 6.48 ±1.50 InI 0.04 ±0.01 0.04 ±0.00 0.04 ±0.00 0.05 ±0.01 0.04 ±0.01 0.04 ±0.01 0.04 ±0.00 0.04 ±0.00 0.04 ±0.00 0.04 ±0.01 CR 0.12 ±0.04 0.15 ±0.03 0.15 ±0.04 0.19 ±0.04 0.16 ±0.04 0.12 ±0.02 0.10 ±0.02 0.11 ±0.03 0.16 ±0.06 0.12 ±0.03 CD 0.28 ±0.08 0.33 ±0.07 0.30 ±0.05 0.30 ±0.05 0.27 ±0.06 0.27 ±0.06 0.21 ±0.04 0.25 ±0.06 0.27 ±0.12 0.21 ±0.08 IcI 1.85 ±0.55 1.66 ±0.50 1.75 ±0.72 2.14 ±0.61 2.34 ±0.68 2.12 ±0.76 3.01 ±0.84 2.55 ±0.61 2.37 ±0.34 3.55 ±1.94 MF 2.11 ±0.52 1.73 ±0.44 1.92 ±0.32 1.53 ±0.41 1.86 ±0.54 2.08 ±0.88 2.10 ±0.41 2.23 ±0.85 4.48 ±1.46 2.50 ±1.08 4458.43 ±142.65 4410.71 ±173.41 4472.01 ±208.27 4462.54 ±209.18 4626.97 ±162.68 4611.94 ±195.25 4603.46 ±130.76 4580.65 ±130.81 4132.84 ±123.98 4666.92 ±145.29 LF HF NM 4069.69 4768.18 698.49 ±123.69 ±164.53 ±72.65 4027.73 4709.60 681.87 ±140.63 ±161.41 ±106.08 4109.14 4795.75 686.61 ±159.39 ±155.80 ±103.87 4054.99 4734.83 679.84 ±164.57 ±192.25 ±96.31 4203.95 5005.74 801.80 ±106.65 ±154.80 ±86.85 4234.59 4890.46 655.88 ±147.38 ±184.53 ±91.30 4198.16 4879.78 681.63 ±104.90 ±147.29 ±103.93 4237.81 4900.20 662.39 ±122.23 ±123.36 ±93.38 3769.95 4341.19 571.24 ±176.51 ±120.80 ±55.71 4274.71 4898.46 623.76 ±84.27 ±111.44 ±116.87 99 Online Resource 2 Locality 1 2 3 4 5 6 7 8 9 10 ND 0.04 ±0.01 0.05 ±0.00 0.05 ±0.00 0.05 ±0.00 0.05 ±0.00 0.05 ±0.00 0.05 ±0.00 0.05 ±0.01 0.04 ±0.00 0.06 ±0.01 IN 0.27 ±0.04 0.30 ±0.03 0.24 ±0.03 0.29 ±0.06 0.31 ±0.07 0.35 ±0.06 0.42 ±0.09 0.41 ±0.04 0.28 ±0.04 0.28 ±0.07 LF 4069.05 ±129.50 4103.23 ±184.94 4051.22 ±181.26 4205.78 ±255.10 3686.44 ±94.83 4387.85 ±324.03 4215.64 ±137.96 4440.11 ±205.10 4198.56 ±279.52 4374.48 ±94.77 HF 4421.53 ±114.81 4487.66 ±127.16 4429.06 ±148.95 4541.89 ±235.64 4007.51 ±139.53 4782.38 ±282.56 4628.13 ±111.98 4829.18 ±187.38 4627.34 ±237.70 4752.86 ±99.53 NM 354.08 ±31.78 365.36 ±63.84 377.81 ±56.02 336.90 ±34.26 342.31 ±36.06 392.55 ±51.10 396.79 ±55.15 385.78 ±44.10 428.79 ±60.98 377.81 ±44.03 PF 4231.73 ±145.24 4302.14 ±153.46 4231.38 ±164.48 4351.32 ±240.51 3863.47 ±96.02 4581.54 ±289.99 4418.89 ±112.00 4627.57 ±187.86 4416.06 ±243.14 4566.55 ±92.95 100 Online Resource 3 Locality 1 2 3 4 5 6 7 8 9 10 SVL 20.25 ±0.99 20.25 ±1.15 19.90 ±0.70 20.53 ±0.78 19.51 ±0.87 19.83 ±1.02 19.81 ±0.61 20.06 ±0.76 21.73 ±0.61 19.06 ±0.51 HL 6.29 ±0.28 6.28 ±0.24 6.05 ±0.25 5.80 ±0.44 5.96 ±0.27 6.02 ±0.31 6.19 ±0.19 5.92 ±0.35 6.46 ±0.15 5.92 ±0.27 HW 6.70 ±0.28 6.84 ±0.26 6.58 ±0.33 6.81 ±0.32 6.49 ±0.30 6.79 ±0.33 6.83 ±0.19 6.66 ±0.24 7.02 ±0.21 6.39 ±0.18 SL 2.86 ±0.16 2.81 ±0.20 2.74 ±0.21 2.63 ±0.21 2.74 ±0.18 2.80 ±0.14 3.01 ±0.22 2.74 ±0.22 2.93 ±0.22 2.77 ±0.23 EN 2.03 ±0.07 1.99 ±0.12 1.92 ±0.07 1.93 ±0.07 1.91 ±0.09 1.95 ±0.11 1.99 ±0.11 1.96 ±0.15 2.00 ±0.07 1.89 ±0.10 IN 2.75 ±0.12 2.81 ±0.18 2.73 ±0.10 2.78 ±0.09 2.70 ±0.10 2.73 ±0.08 2.72 ±0.09 2.75 ±0.10 2.90 ±0.09 2.61 ±0.09 EL 2.56 ±0.14 2.60 ±0.14 2.63 ±0.14 2.70 ±0.15 2.55 ±0.09 2.61 ±0.22 2.58 ±0.14 2.63 ±0.14 2.51 ±0.13 2.62 ±0.15 IO 5.50 ±0.29 5.70 ±0.13 5.57 ±0.21 5.66 ±0.26 5.46 ±0.17 5.54 ±0.23 5.56 ±0.16 5.78 ±0.16 6.02 ±0.13 5.35 ±0.19 TYM 1.14 ±0.12 1.03 ±0.12 1.03 ±0.11 1.01 ±0.14 1.10 ±0.11 1.05 ±0.15 1.08 ±0.15 0.95 ±0.12 1.16 ±0.09 1.05 ±0.09 FAL 4.67 ±0.29 4.63 ±0.31 4.51 ±0.18 4.69 ±0.19 4.53 ±0.18 4.40 ±0.23 4.51 ±0.28 4.42 ±0.20 4.91 ±0.33 4.36 ±0.16 UAL 5.39 ±0.24 5.32 ±0.21 5.11 ±0.22 4.99 ±0.42 5.04 ±0.20 5.15 ±0.21 5.05 ±0.32 5.18 ±0.34 5.21 ±0.43 4.72 ±0.42 HAND1 3.00 ±0.25 3.03 ±0.31 2.68 ±0.29 2.92 ±0.24 2.78 ±0.26 3.00 ±0.26 2.98 ±0.26 3.09 ±0.19 3.19 ±0.27 2.98 ±0.18 HAND2 2.72 ±0.21 2.77 ±0.23 2.60 ±0.25 2.73 ±0.24 2.59 ±0.15 2.70 ±0.19 2.63 ±0.28 2.73 ±0.26 3.00 ±0.17 2.65 ±0.16 HAND3 3.86 ±0.24 3.86 ±0.27 3.65 ±0.26 3.95 ±0.31 3.83 ±0.20 3.91 ±0.23 3.94 ±0.24 3.94 ±0.27 4.23 ±0.19 3.91 ±0.19 HAND4 2.34 ±0.13 2.50 ±0.17 2.34 ±0.09 2.50 ±0.15 2.36 ±0.16 2.49 ±0.18 2.42 ±0.15 2.38 ±0.23 2.67 ±0.50 2.44 ±0.19 WFD 0.55 ±0.05 0.57 ±0.09 0.52 ±0.07 0.59 ±0.06 0.51 ±0.05 0.59 ±0.08 0.63 ±0.08 0.64 ±0.08 0.59 ±0.08 0.60 ±0.06 TL 9.26 ±0.37 9.53 ±0.33 9.19 ±0.25 9.20 ±0.50 8.99 ±0.26 9.17 ±0.51 9.06 ±0.33 9.03 ±0.43 10.11 ±0.49 9.03 ±0.21 FL 7.81 ±0.38 7.83 ±0.43 7.58 ±0.41 8.07 ±0.31 7.68 ±0.34 7.47 ±0.34 7.56 ±0.26 7.83 ±0.41 8.31 ±0.55 7.58 ±0.41 LL 8.48 ±0.44 9.03 ±0.46 8.40 ±0.32 8.49 ±0.56 8.55 ±0.26 8.54 ±0.40 8.48 ±0.38 8.73 ±0.47 9.38 ±0.46 8.34 ±0.49 DPT 0.71 ±0.07 0.78 ±0.08 0.77 ±0.10 0.75 ±0.12 0.78 ±0.08 0.74 ±0.09 0.75 ±0.07 0.75 ±0.08 0.81 ±0.08 0.72 ±0.09 WTT 0.63 ±0.12 0.68 ±0.08 0.64 ±0.07 0.63 ±0.13 0.61 ±0.13 0.62 ±0.07 0.65 ±0.05 0.58 ±0.08 0.61 ±0.11 0.65 ±0.11 WTD 0.65 ±0.10 0.69 ±0.10 0.65 ±0.11 0.69 ±0.11 0.66 ±0.08 0.72 ±0.11 0.73 ±0.11 0.80 ±0.11 0.64 ±0.13 0.70 ±0.07 WPF 0.54 ±0.05 0.58 ±0.06 0.55 ±0.05 0.45 ±0.05 0.52 ±0.07 0.55 ±0.05 0.59 ±0.05 0.58 ±0.06 0.50 ±0.09 0.52 ±0.04 101 Online Resource 4 Locality 1 2 3 4 5 6 7 8 9 10 SVL 19.74 ±0.62 19.25 ±0.63 19.76 ±0.71 18.75 ±0.46 20.63 ±0.10 18.58 ±0.69 19.27 ±0.35 18.02 ±0.65 19.32 ±0.38 20.46 ±0.70 HL 7.17 ±0.40 6.67 ±0.42 6.90 ±0.41 6.63 ±0.34 7.09 ±0.08 6.50 ±0.38 6.53 ±0.51 6.26 ±0.34 6.57 ±0.39 7.02 ±0.48 HW 7.04 ±0.25 6.68 ±0.19 6.96 ±0.23 6.54 ±0.19 7.26 ±0.19 6.49 ±0.30 6.69 ±0.35 6.23 ±0.24 6.71 ±0.34 7.16 ±0.45 SL 2.87 ±0.27 2.64 ±0.12 2.72 ±0.12 2.86 ±0.14 2.89 ±0.24 2.63 ±0.23 2.71 ±0.17 2.57 ±0.24 2.66 ±0.17 2.25 ±0.21 EN 2.07 ±0.18 1.89 ±0.15 2.03 ±0.05 1.98 ±0.09 2.08 ±0.07 1.84 ±0.15 1.90 ±0.13 1.82 ±0.14 1.91 ±0.07 1.53 ±0.13 IN 2.67 ±0.07 2.49 ±0.14 2.60 ±0.12 2.58 ±0.23 2.80 ±0.31 2.54 ±0.12 2.64 ±0.16 2.44 ±0.12 2.60 ±0.16 2.72 ±0.15 EL 2.74 ±0.16 2.64 ±0.17 2.60 ±0.11 2.66 ±0.12 2.74 ±0.33 2.64 ±0.18 2.63 ±0.11 2.63 ±0.08 2.73 ±0.16 2.55 ±0.14 IO 2.44 ±0.28 2.22 ±0.19 2.33 ±0.14 2.25 ±0.19 2.50 ±0.20 2.12 ±0.14 2.33 ±0.16 2.22 ±0.12 2.42 ±0.19 2.34 ±0.24 TYM 1.05 ±0.14 1.05 ±0.08 1.15 ±0.13 1.00 ±0.12 1.22 ±0.18 1.06 ±0.15 1.01 ±0.10 1.02 ±0.10 1.03 ±0.17 1.02 ±0.14 FAL 4.18 ±0.24 4.19 ±0.17 4.07 ±0.17 4.15 ±0.17 4.36 ±0.25 4.22 ±0.27 4.21 ±0.27 4.25 ±0.25 4.45 ±0.20 4.35 ±0.18 UAL 4.91 ±0.21 4.88 ±0.20 4.95 ±0.25 4.75 ±0.30 5.13 ±0.16 4.75 ±0.27 4.86 ±0.25 4.77 ±0.33 5.04 ±0.25 5.12 ±0.20 HAND1 3.31 ±0.15 3.36 ±0.13 3.33 ±0.17 3.32 ±0.18 3.57 ±0.10 3.33 ±0.19 3.47 ±0.19 3.33 ±0.19 3.58 ±0.19 3.69 ±0.15 HAND2 3.13 ±0.12 3.10 ±0.12 3.17 ±0.15 3.10 ±0.19 3.20 ±0.40 2.91 ±0.11 3.05 ±0.12 2.95 ±0.18 3.16 ±0.11 3.39 ±0.19 HAND3 4.42 ±0.13 4.31 ±0.14 4.49 ±0.14 4.28 ±0.22 4.47 ±0.23 4.02 ±0.18 4.18 ±0.23 4.00 ±0.17 4.21 ±0.17 4.58 ±0.26 HAND4 2.98 ±0.18 2.88 ±0.17 2.98 ±0.20 2.88 ±0.15 2.95 ±0.37 2.71 ±0.21 2.79 ±0.16 2.62 ±0.16 2.82 ±0.15 3.06 ±0.17 WFD 0.50 ±0.05 0.46 ±0.08 0.50 ±0.07 0.47 ±0.09 0.54 ±0.06 0.47 ±0.06 0.49 ±0.05 0.42 ±0.04 0.45 ±0.05 0.60 ±0.05 TL 8.88 ±0.38 8.69 ±0.31 8.78 ±0.20 8.63 ±0.33 8.90 ±0.08 8.36 ±0.24 8.79 ±0.39 8.42 ±0.51 8.83 ±0.33 9.05 ±0.33 FL 7.65 ±0.34 7.68 ±0.27 7.92 ±0.30 7.71 ±0.34 7.88 ±0.07 7.40 ±0.31 7.85 ±0.29 7.53 ±0.34 7.89 ±0.36 7.82 ±0.42 LL 8.77 ±0.35 8.90 ±0.13 8.87 ±0.31 8.71 ±0.34 9.16 ±0.05 8.66 ±0.35 9.00 ±0.41 8.60 ±0.41 9.06 ±0.32 8.96 ±0.48 DPT 0.63 ±0.06 0.64 ±0.05 0.62 ±0.08 0.65 ±0.06 0.66 ±0.10 0.60 ±0.04 0.68 ±0.07 0.59 ±0.06 0.68 ±0.09 0.64 ±0.04 WTT 0.48 ±0.06 0.43 ±0.06 0.43 ±0.06 0.43 ±0.04 0.41 ±0.08 0.37 ±0.04 0.36 ±0.08 0.36 ±0.05 0.39 ±0.06 0.45 ±0.07 WTD 0.65 ±0.05 0.69 ±0.09 0.68 ±0.06 0.68 ±0.08 0.72 ±0.19 0.62 ±0.06 0.64 ±0.06 0.55 ±0.06 0.59 ±0.07 0.79 ±0.05 WPF 0.40 ±0.04 0.36 ±0.06 0.36 ±0.05 0.38 ±0.05 0.40 ±0.24 0.35 ±0.03 0.40 ±0.04 0.31 ±0.02 0.33 ±0.04 0.46 ±0.04 102 Online Resource 5 Locality paleovarzensis system Haplotype 1 2 H01 7 1 H02 1 H03 6 H04 2 H05 1 H06 7 H07 4 H08 1 H09 1 H10 H11 H12 H13 H14 H15 H16 H17 H18 H19 H20 H21 H22 H23 H24 H25 H26 H27 H28 H29 nidicola-masniger system H01 11 1 H02 5 H03 6 H04 H05 H06 H07 H08 H09 H10 H11 H12 H13 H14 H15 H16 H17 H18 H19 H20 H21 H22 H23 H24 H25 H26 - 3 13 1 1 - 4 14 1 - 5 14 3 - 6 12 1 - 7 3 1 2 2 1 - 8 10 1 1 - 9 12 2 1 1 2 - 10 7 3 1 1 Collection # INPA-H 29176 APL 12625 INPA-H 29071 APL 12631 APL 12632 APL 954 INPA-H 29185 INPA-H 29145 INPA-H 29184 INPA-H 29117 INPA-H 29114 INPA-H 29124 APL 12681 INPA-H 29135 INPA-H 29171 INPA-H 29193 INPA-H 29188 INPA-H 29194 INPA-H 29195 INPA-H 29146 INPA-H 29155 INPA-H 29079 INPA-H 29091 INPA-H 29098 INPA-H 29106 INPA-H 29097 INPA-H 29086 INPA-H 29082 INPA-H 29083 GenBank # JQ966835 JQ966836 JQ966837 JQ966838 JQ966839 JQ966840 JQ966841 JQ966842 JQ966843 JQ966844 JQ966845 JQ966846 JQ966847 JQ966848 JQ966849 JQ966850 JQ966851 JQ966852 JQ966853 JQ966854 JQ966855 JQ966856 JQ966857 JQ966858 JQ966859 JQ966860 JQ966861 JQ966862 JQ966863 13 1 - 7 1 2 2 1 - 9 5 - 2 12 1 - 1 13 1 - 9 1 1 1 - 2 10 1 1 - 5 8 1 INPA-H 28140 INPA-H 28173 INPA-H 28145 INPA-H 28130 INPA-H 28134 INPA-H 28142 INPA-H 28177 APL 14320 APL 14322 INPA-H 28092 INPA-H 28070 APL 12763 APL 14336 APL 14399 APL 14397 INPA-H 28079 APL 14404 APL 14406 INPA-H 28102 APL 14252 INPA-H 28093 APL 14251 APL 14250 APL 12957 APL 12960 APL 12958 JQ966864 JQ966865 JQ966866 JQ966867 JQ966868 JQ966869 JQ966870 JQ966871 JQ966872 JQ966873 JQ966874 JQ966875 JQ966876 JQ966877 JQ966878 JQ966879 JQ966880 JQ966881 JQ966882 JQ966883 JQ966884 JQ966885 JQ966886 JQ966887 JQ966888 JQ966889 103 Online Resource 6 104 Online Resource 7 C1 C1 C2 C3 C4 Total 27 5 4 1 37 C1 C2 C3 C4 Total 22 6 3 1 32 C1 C2 C3 C4 Total 21 8 5 0 34 Genetic cluster C2 C3 C4 Temperature-adjusted call traits 11 5 10 9 0 3 2 5 0 0 0 0 22 10 13 Temperature- and body-size-adjusted call traits 15 4 12 7 2 2 2 6 0 0 0 0 24 12 14 Morphology 16 15 1 6 3 0 5 7 1 0 1 0 27 26 2 % correct 51 53 45 0 50 42 41 55 0 43 40 35 39 0 38 105 Online Resource 8 C1 C2 C1 C2 C3 C4 C5 C6 C7 Total 5 2 1 2 0 0 0 10 1 25 0 0 2 0 3 31 C1 C2 C3 C4 C5 C6 C7 Total 3 2 5 1 0 0 0 11 1 23 0 1 2 0 2 29 C1 C2 C3 C4 C5 C6 C7 Total 8 1 0 0 0 0 0 9 0 19 2 1 2 0 1 25 Genetic cluster C3 C4 C5 C6 Temperature-adjusted call traits 0 1 0 0 2 1 2 0 9 0 0 0 0 1 3 3 0 2 3 2 0 2 1 5 0 0 0 0 11 7 9 10 Temperature- and body-size-adjusted call traits 2 1 0 0 4 3 0 0 4 0 0 1 1 1 2 2 0 2 2 3 0 2 2 4 0 0 0 0 11 9 6 10 Morphology 0 0 0 0 6 3 2 4 6 1 0 0 2 0 4 2 1 2 0 2 0 2 0 3 1 2 2 1 16 10 8 12 C7 % correct 1 6 0 1 1 0 7 16 63 66 90 10 30 63 70 59 1 6 0 2 1 0 8 18 38 61 40 10 20 50 80 48 0 3 1 1 3 3 3 14 100 50 60 0 0 38 30 41 106 SÍNTESE Em conjunto, os resultados dos três capítulos aqui apresentados ressaltam a complementaridade de respostas proporcionada pela investigação de distintas classes de caracteres e de diferentes níveis da hierarquia biológica no estudo de questões em biogeografia histórica. Os dois primeiros capítulos transitaram do nível intraindividual ao interespecífico no estudo de como caracteres comportamentais importantes em escolha e reconhecimento sexual variam no espaço. Em adição, foi possível determinar como fatores individuais e ambientais afetam esses sinais. No terceiro capítulo, foi possível realizar uma abordagem comparativa para avaliar a resposta de diferentes classes de marcadores à história biogeográfica das espécies. Por meio da amostragem realizada, verificou-se que Allobates paleovarzensis ocorre em ambas as margens do Rio Amazonas. Sua área de distribuição é maior do que aquela anteriormente conhecida, estendendo-se até o município de Codajás, a oeste, e até o município de Barcelos, ao norte. Nenhum dos caracteres acústicos analisados foi indicado como potencial sinal para reconhecimento social ou seleção sexual na espécie estudada, e futuras investigações envolvendo experimentos de reconhecimento sexual intraespecífico deverão elucidar as potenciais implicações do padrão observado. Todas as propriedades estáticas (no caso, espectrais) do canto de anúncio estiveram sob forte regulação morfológica. Portanto, esses caracteres podem ser indicadores honestos dos seus fatores controladores. Essas propriedades também deverão estar sujeitas à seleção estabilizadora por parte das fêmeas da espécie. Todas as propriedades acústicas temporais relacionadas às notas estiveram correlacionadas com a temperatura ambiental. Uma vez que caracteres relacionados ao canto (conjunto de notas) não variaram de acordo com os fatores individuais e ambientais avaliados, acredita-se que estes sejam sujeitos à regulação social. Tanto caracteres espectrais quanto temporais diferiram entre lados opostos do Rio Amazonas. Todavia, lado do rio e distância geográfica não afetaram significativamente a variabilidade acústica quando todos os parâmetros acústicos foram sumarizados em componentes principais. Esses resultados indicam que forças evolutivas estabilizadoras locais devem ser importantes no processo de diferenciação do sinal sexual de A. paleovarzensis. Com relação ao sistema de estudo composto por populações de Allobates nidicola e A. masniger, propriedades espectrais do canto de anúncio foram mais distintivas que propriedades temporais entre populações e entre espécies, com maior distinção ao nível 107 interpopulacional quando comparado ao nível interespecífico. O entendimento do rio como barreira levando à especiação no sistema de estudo seria beneficiado por abordagens que avaliem a variabilidade dos caracteres acústicos ao nível intraindividual, bem como o papel de propriedades espectrais no reconhecimento e escolha sexual por parte de indivíduos de ambas as margens. Houve acentuada plasticidade em caracteres temporais do canto de anúncio, os quais tiveram sua variabilidade fortemente induzida pela temperatura ambiental. Caracteres espectrais também mostraram correlação positiva, embora moderada, com essa variável ambiental. Os efeitos do tamanho corporal foram restritos às propriedades espectrais do canto de anúncio. As propriedades modulação da nota e intervalo entre notas foram pouco afetadas pela temperatura ambiental e tamanho corporal, sugerindo que outros fatores subjacentes devem estar relacionados à sua regulação. O efeito do Rio Madeira como barreira foi significativo entre todas as variáveis acústicas analisadas, mesmo após a inclusão de tamanho corporal como covariável. Foi detectada uma interação entre geografia (localidade) e tamanho corporal na diferenciação de caracteres espectrais do sinal sexual. A ausência de efeitos de isolamento por distância indica que pressões seletivas estabilizadoras locais devem ser mais importantes que a deriva genética na evolução da diferenciação do canto de anúncio no sistema nidicola/masniger. Observou-se significativa estruturação genética populacional em ambos os sistemas de estudo. Tal padrão pode estar relacionado ao pequeno tamanho corporal e à territorialidade das espécies. A estrutura genética dos dois sistemas de estudo exibiu assinaturas congruentes com a transposição de grandes rios amazônicos. A efetividade do baixo Rio Madeira como uma barreira ao fluxo gênico foi maior do que a do médio Amazonas. Entretanto, características idiossincráticas dos sistemas de amostragem e das espécies estudadas podem ter contribuído para o resultado observado. Por exemplo, as espécies de desenvolvimento larval direto (A. nidicola e A. masniger) apresentaram estruturação genética populacional mais expressiva, bem como um efeito mais pronunciado do rio como barreira em relação à A. paleovarzensis, cujos girinos completam seu desenvolvimento em corpos d’água. Allobates nidicola e A. masniger podem ser consideradas espécies distintas sob os conceitos evolutivo e filogenético de espécie. Populações de Allobates paleovarzensis pertencentes à interflúvios distintos parecem não ter sido sujeitas ao nível e à duração do isolamento genético necessários para que a diferenciação populacional atinja os critérios de especiação sob os conceitos de espécie baseados em monofilia e fenótipo. Além dos efeitos observados de rios como barreiras biogeográficas, uma grande porção da diferenciação genética foi correlacionada com 108 distância geográfica linear, evidenciando a importante atuação da deriva genética na diversificação do marcador mitocondrial estudado. Os caracteres acústicos e morfológicos não responderam às distâncias lineares entre populações. Em adição, análises de função discriminante falharam em classificar corretamente os indivíduos analisados em seus agrupamentos genéticos de acordo com seus fenótipos. Esse resultado é surpreendente levando-se em consideração a alta estruturação genética observada, e sugere que limitações evolutivas, as quais não agiram sobre marcadores moleculares neutros, podem estar impedindo a diferenciação fenotípica. 109 REFERÊNCIAS BIBLIOGRÁFICAS Adams, D.C.; Berns, C.M.; Kozak, K H.; Wiens, J. 2009. Are rates of species diversification correlated with rates of morphological evolution? Proceedings of the Royal Society B, 276: 2729-2738. Amézquita, A.; Hödl, W.; Lima, A.P.; Castellanos, L.; Erdtmann, L.; Araújo, M.C. 2006. Masking interference and the evolution of the acoustic communication system in the Amazonian dendrobatid frog Allobates femoralis. Evolution, 60: 1874-1887. Amézquita, A.; Lima, A.P.; Jehle, R.; Castellanos, L.; Ramos, O.; Crawford, A.J.; Gasser, H.; Hödl, W. 2009. Calls, colours, shapes, and genes: A multi-trait approach to the study of geographic variation in the Amazonian frog Allobates femoralis. Biological Journal of the Linnean Society, 98: 826-838. Anderson, M.J.; Legendre, P. 1999. An empirical comparison of permutation methods for tests of partial regression coefficients in a linear model. Journal of Statistical Computation and Simulation, 62: 271-303. Angulo, A.; Icochea, J. 2010. Cryptic species complexes, widespread species and conservation: lessons from Amazonian frogs of the Leptodactylus marmoratus group (Anura: Leptodactylidae). Systematics and Biodiversity, 8: 357-370. Angulo, A.; Reichle, S. 2008. Acoustic signals, species diagnosis, and species concepts: the case of a new cryptic species of Leptodactylus (Amphibia, Anura, Leptodactylidae) from the Chapare region, Bolivia. Zoological Journal of the Linnean Society, 152: 58-77. Antonelli, A.; Quijada-Mascareñas, A.; Crawford, A.J.; Bates, J.M.; Velazco, P.M; Wüster, W. 2010. Molecular studies and phylogeography of Amazonian tetrapods and their relation to geological and climatic models. In: Hoorn, C.; Wesselingh, F.P. (Eds). Amazonia, landscape and species evolution. Blackwell Publishing, Oxford. p. 386-403. Avise, J.C.; Arnold, J.; Ball, R.M.; Bermingham, E.; Lamb, T.; Neigel, J.E.; Reeb, C.A.; Saunders, N.C. 1987. Intraspecific phylogeography: the mitochondrial DNA bridge between population genetics and systematics. Annual Review of Ecology and Systematics, 18: 489-522. 110 Avise, J.C. 2000. Phylogeography: the history and formation of species. Harvard University Press, Cambridge. Avise, J.C. 2004. Molecular Markers, Natural History and Evolution. Sinauer Associates Inc., Sunderland. Avise, J.C. 2009. Phylogeography: retrospect and prospect. Journal of Biogeography, 36: 3-15. Ayres, J.M.; Clutton-Brock, T.H. 1992. River boundaries and species range size in Amazonian primates. The American Naturalist, 140: 531-537. Bee, M.A.; Cook, J.M.; Love, E.K.; O’Bryan, L.R.; Pettitt, B.A.; Schrode, K.; Vélez, A. 2010. Assessing acoustic signal variability and the potential for sexual selection and social recognition in boreal chorus frogs (Pseudacris maculata). Ethology, 116: 564-576. Bee, M.A.; Perrill, S.A.; Owen, P.C. 2000. Male green frogs lower the pitch of acoustic signals in defense of territories: a possible dishonest signal of size? Behavioral Ecology, 11: 169-177. Behr, O.; von Helversen, O.; Heckel, G.; Nagy, M.; Voigt, C.C.; Mayer, F. 2007. Territorial songs indicate male quality in the sac-winged bat Saccopteryx bilineata (Chiroptera, Emballonuridae). Behavioral Ecology, 17: 810-817. Blomberg, S.P.; Garland Jr., T.; Ives, A.R. 2003. Testing for phylogenetic signal in comparative data: behavioral traits are more labile. Evolution, 57: 717-745. Bonnet, E.; Van de Peer, Y. 2002. ZT: a software tool for simple and partial Mantel tests. Journal of Statistical Software, 7: 1-12. Boul, K.E.; Funk, W.C.; Darst, C.R.; Cannatella, D.C.; Ryan, M.J. 2007. Sexual selection drives speciation in an Amazonian frog. Proceedings of the Royal Society B, 274: 399-406. Buckley, D. 2009. Toward an organismal, integrative, and iterative phylogeography. Journal of Biogeography, 36: 3-15. Caldwell, J.P.; Lima, A.P. 2003. A new Amazonian species of Colostethus (Anura: Dendrobatidae) with a nidicolous tadpole. Herpetologica, 59: 219-234. Camargo, A.; de Sá, R.; Heyer, W.R. 2006. Phylogenetic analyses of mtDNA sequences reveal three cryptic lineages in the widespread neotropical frog Leptodactylus fuscus (Schneider, 1799) (Anura, Leptodactylidae). Biological Journal of the Linnean Society, 87: 325-341. 111 Campbell, P.; Pasch, B.; Pino, J.L.; Crino, O.L.; Phillips, M.; Phelps, S.M. 2010. Geographic variation in the songs of Neotropical singing mice: testing the relative importance of drift and local adaptation. Evolution, 64: 1955-1972. Capparella, A.P. 1988. Genetic variation in Neotropical birds: implications for the speciation process. In: Ouellet, H. (Ed). Acta XIX Congressus Internationalis Ornithologici. University of Ottawa Press, Ottaxa. p. 1658-1664. Castellano, S.; Cuatto, B.; Rinella, R.; Rosso, A.; Giacoma, C. 2002. The advertisement call of the European treefrogs (Hyla arborea): a multilevel study of variation. Ethology, 108: 75-89. Castellano, S.; Giacoma, C. 1998. Stabilizing and directional female choice for male calls in the European green toads. Animal Behaviour, 56: 275-287. Castellano, S.; Giacoma, C.; Dujsebayeva, T. 2000. Morphometric and advertisement call geographic variation in polyploid green toads. Biological Journal of the Linnean Society, 70: 341-360. Chakraborty, M.; Mangiamele, L.A.; Burmeister, S.S. 2010. Neural activity patterns in response to interspecific and intraspecific variation in mating calls in the Túngara Frog. PLoS One, 5: e12898. Charif, R.A.; Clark, C.W.; Fristrup, K.M. 2004. Raven 1.2 User’s Manual. Cornell Laboratory of Ornithology, Ithaca, New York. Charif, R.A.; Waack, A.M.; Strickman, L.M. 2008. Raven Pro 1.3 User’s Manual. Cornell Laboratory of Ornithology, Ithaca, New York. Cherry, L.M.; Case, S.M.; Wilson, A.C. 1978. Frog perspective on the morphological difference between humans and chimpanzees. Science, 200: 209-211. Clement, M.; Posada, D.; Crandall, K.A. 2000. TCS: a computer program to estimate gene genealogies. Molecular Ecology, 9: 1657-1659. Cohn-Haft, M. 2000. A case study in amazonian biogeography: Vocal and DNA sequence variation in Hemitriccus flycatchers. Tese de Doutorado, Luisiana State University, Baton Rouge. 112 Colwell, R.K. 2000. A barrier runs through it… or maybe just a river. Proceedings of the National Academy of Sciences, 97: 13470-13472. Corander, J.; Marttinen, P.; Sirén J.; Tang, J. 2008. Enhanced Bayesian modelling in BAPS software for learning genetic structures of populations. BMC Bioinformatics, 9: 539. Costa, J.B.S.; Bemerguy, R.L.; Hasui, Y.; Borges, M.S. 2001. Tectonics and paleogeography along the Amazon River. Journal of South American Earth Sciences, 14: 335-347. Coyne, J.A.; Orr, H.A. 2004. Speciation. Sinauer Associates Inc., Sunderland. Crawford, A.J. 2003. Huge populations and old species of Costa Rican and Panamanian dirt frogs inferred from mitochondrial and nuclear gene sequences. Molecular Ecology, 12: 2525-2540. Crawford, A.J.; Lips, K.R.; Bermingham, E. 2010. Epidemic disease decimates amphibian abundance, species diversity, and evolutionary history in the highlands of central Panama. Proceedings of the National Academy of Sciences, 107: 13777-13782. Creer, S.; Thorpe, R.S.; Malhotra, A.; Chou, W.H.; Stenson, A.G. 2004. The utility of AFLPs for supporting mitochondrial DNA phylogeographical analyses in the Taiwanese bamboo viper, Trimeresurus stejnegeri. Journal of Evolutionary Biology, 17: 100-107. Crisci, J.V.; Katinas, L.; Posadas, P. 2003. Historical biogeography: An Introduction. Harvard University Press, Cambridge. Cronquist, A. 1978. Once again, what is a species? In: Knutson, L.V. (Ed). Biosystematics in Agriculture. Allenheld Osmun, Montclair, New Jersey. p. 3-20. Dall, S.R.X. 1997. Behaviour and speciation. Trends in Ecology and Evolution, 12: 209-210. Dayrat, B. 2005. Towards integrative taxonomy. Biological Journal of the Linnean Society, 85: 407-415. de Queiroz, K. 2007. Species concepts and species delineation. Systematic Biology, 56: 879-886. Diniz-Filho, J.A.F.; Telles, M.P.C.; Bonatto, S.L.; Eizirik, E.; Freitas, T.R.O.; de Marco, P.; Santos, F.R.; Solé-Cava, A.M.; Soares, T.N. 2008. Mapping the evolutionary twilight zone: molecular markers, populations and geography. Journal of Biogeography, 35: 753-763. 113 Drummond, A.J.; Rambaut, A. 2007. BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evolutionary Biology, 7: 214. Elmer, K.R.; Dávila, J.A.; Lougheed, S.C. 2007. Cryptic diversity and deep divergence in an upper Amazonian leaflitter frog, Eleutherodactylus ockendeni. BMC Evolutionary Biology, 7: 247. Emerson, K.J.; Merz, C.R.; Catchen, J.M.; Hohenlohe, P.A.; Cresko, W.A.; Bradshaw, W.E.; Holzapfel, C.M. 2010. Resolving postglacial phylogeography using high-throughput sequencing. Proceedings of the National Academy of Sciences, 107: 16196-16200. Erdtmann, L.; Amézquita, A. 2009. Differential evolution of advertisement call traits in DartPoison Frogs (Anura: Dendrobatidae). Ethology, 115: 801-811. Excoffier, L.; Lischer, H.E.L. 2010. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Molecular Ecology Research, 10: 564-567. Excoffier, L.; Smouse, P.; Quattro, J. 1992. Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics, 131: 479-491. Fernandes, A.M.; Wink, M.; Aleixo, A. 2012. Phylogeography of the chestnut tailed antbird (Myrmeciza hemimelaena) clarifies the role of rivers in Amazonian biogeography. Journal of Biogeography, doi:10.1111/j.1365-2699.2012.02712.x Figueiredo, J.; Hoorn, C.; van der Ven, P.; Soares, E. 2009. Late Miocene onset of the Amazon River and the Amazon deep-sea fan: Evidence from the Foz do Amazonas Basin. Geology, 37: 619-622. Fischer, J.; Kitchen, D.M.; Seyfarth, R.M.; Cheney, D.L. 2004. Baboon loud calls advertise male quality: acoustic features and their relation to rank, age, and exhaustion. Behavioral Ecology and Sociobiology, 56: 140-148. Fouquet, A.; Gilles, A.; Vences, M.; Marty, C.; Blanc, M.; Gemmell, N.J. 2007. Underestimation of species richness in Neotropical frogs revealed by mtDNA analyses. PLoS One, 32: e1109. 114 Fouquet, A.; Noonan, B.P.; Rodrigues, M.T.; Pech, N.; Gilles, A.; Gemmell, N.J. Multiple Quaternary Refugia in the Eastern Guiana Shield Revealed by Comparative Phylogeography of 12 Frog Species. Systematic Biology, in press. Fouquet, A.; Recoder, R.; Teixeira–Jr, M.; Cassimiro, J.; Amaro, R.C.; Camacho, A.; Damasceno, R.; Carnaval, A.C.; Moritz, C.; Rodrigues, M.T. 2012. Molecular phylogeny and morphometric analyses reveal deep divergence between Amazonia and Atlantic Forest species of Dendrophryniscus. Molecular Phylogenetics and Evolution, 62: 826-838. Friedl, T.W.P.; Klump, G.M. 2002. The vocal behaviour of male European treefrogs (Hyla arborea): implications for inter- and intrasexual selection. Behaviour, 139: 113-136. Fu, Y.X. 1997. Statistical tests of neutrality of mutations against population growth, hitchhiking and background selection. Genetics, 147: 915-925. Funk, D.J.; Omland, K.E. 2003. Species-level paraphyly and polyphyly: frequency, causes, and consequences, with insights from animal mitochondrial DNA. Annual Review of Ecology, Evolution, and Systematics, 34: 397-423. Funk, W.C.; Caldwell, J.P.; Peden, C.E.; Padial, J.M.; de la Riva, I.; Cannatella, D.C. 2007. Tests of biogeographic hypotheses for diversification in the Amazonian forest frog, Physalaemus petersi. Molecular Phylogenetics and Evolution, 44: 825-837. Funk, W.C.; Caminer, M.; Ron, S.R. 2012. High levels of cryptic species diversity uncovered in Amazonian frogs. Proceedings of the Royal Society B, 279: 1806-1814. Gascon, C.; Lougheed, S.C.; Bogart, J.P. 1996. Genetic and morphological variation in Vanzolinius discodactylus: a direct test of the riverine barrier hypothesis. Biotropica 28, 376387. Gascon, C.; Lougheed, S.C.; Bogart, J. P. 1998. Patterns of genetic population differentiation in four species of Amazonian frogs: A test of the Riverine Barrier Hypothesis. Biotropica, 30: 104-119. Gascon, C.; Malcolm, J.R.; Patton, J.L.; Silva, M.N.F.; Bogarti, J.P.; Lougheed, S.P.; Peres, C.A.; Neckel, S.; Boag, P.T. 2000. Riverine barriers and the geographic distribution of Amazonian species. Proceedings of the National Academy of Sciences, 97: 13672-13677. 115 Gasser, H.; Amézquita, A.; Hödl, W. 2009. Who is calling? Intraspecific call variation in the aromobatid frog Allobates femoralis. Ethology, 115: 596-607. Gerhardt, H.C. 1991. Female mate choice in treefrogs: static and dynamic acoustic criteria. Animal Behaviour, 42: 615-635. Gerhardt, H.C.; Huber, F. 2002. Acoustic communication in insects and anurans: common problems and diverse solutions. University of Chicago Press, Chicago. Goicochea, N.; De La Riva, I.; Padial, J.M. 2009. Recovering phylogenetic signal from frog mating calls. Zoologica Scripta, 39: 141-154. Gomez, D.; Richardson, C.; Théry, M.; Lengagne, T.; Léna, J.; Plénet, S.; Joly, P. 2011. Multimodal signals in male European treefrog (Hyla arborea) and the influence of population isolation on signal expression. Biological Journal of the Linnean Society, 103: 633-647. Grant, T.; Frost, D.R.; Caldwell, J.P.; Gagliardo, R.; Haddad, C.F.B.; Kok, P.J.R.; Means, B.D.; Noonan, B.P.; Schargel, W.; Wheeler, W.C. 2006. Phylogenetic systematics of dart-poison frogs and their relatives (Anura: Athesphatanura: Dendrobatidae). Bulletin of the American Museum of Natural History, 299: 1-262. Groeneveld, L.F.; Weisrock, D.W.; Rasoloarison, R.M.; Yoder, A.D.; Kappeler, P.M. 2009. Species delimitation in lemurs: multiple genetic loci reveal low levels of species diversity in the genus Cheirogaleus. BMC Evolutionary Biology, 9: 30. Guerra, M.A.; Ron, S.R. 2008. Mate choice and courtship signal differentiation promotes speciation in an Amazonian frog. Behavioral Ecology, 19: 1128-1135. Guillot, G.; Renaud, S.; Ledevin, R.; Michaux, J.; Claude, J. 2012. A unifying model for the analysis of phenotypic, genetic and geographic data. Systematic Biology, doi: 10.1093/sysbio/sys038. Haffer, J. 1969. Speciation in Amazonian forest birds. Science, 165: 131-137. Hall, T.A. 1999. BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41: 95-98. 116 Hansen, T.F.; Orzack, S.H. 2005. Assessing current adaptation and phylogenetic inertia as explanations of trait evolution: the need for controlled comparisons. Evolution, 59: 20632072. Harpending, R.C. 1994. Signature of ancient population growth in a low-resolution mitochondrial DNA mismatch distribution. Human Biology, 66: 591-600. Hayes, F.E.; Sewlal, J.N. 2004. The Amazon River as a dispersal barrier to passerine birds: effects of river width, habitat and taxonomy. Journal of Biogeography, 31: 1809-1818. Heyer, R.W.; Reid, Y.R. 2003. Does advertisement call variation coincide with genetic variation in the genetically diverse frog taxon currently known as Leptodactylus fuscus (Amphibia: Leptodactylidae)? Anais da Academia Brasileira de Ciências, 75: 39-54. Hoorn, C. 1994. An environmental reconstruction of the palaeo-Amazon River system (Middle– Late Miocene, NW Amazonia). Palaeogeography, Palaeoclimatology, Palaeoecology, 112: 187-238. Hoorn, C.; Wesselingh, F.P.; ter Steege, H.; Bermudez, M.A.; Mora, A.; Sevink, J.; Sanmartín, I.; Sanchez-Meseguer, A.; Anderson, C.L.; Figueiredo, J.P.; Jaramillo, C.; Riff, D.; Negri, F.R.; Hooghiemstra, H.; Lundberg, J.; Stadler, T.; Särkinen, T.; Antonelli, A. 2010. Amazonia through time: Andean uplift, climate change, landscape evolution, and biodiversity. Science, 330: 927-931. Hubert, N.; Meyer, C.P.; Bruggemann, H.J.; Guérin, F.; Komeno, R.J.L.; Espiau, B.; Causse, R.; Williams, J.T.; Planes, S. 2012. Cryptic diversity in Indo-Pacific coral-reef fishes revealed by DNA-barcoding provides new support to the centre-of-overlap hypothesis. PloS One, 7: e28987. Hutchison, D.W.; Templeton, A.R. 1999. Correlation of pairwise genetic and geographic distance measures: inferring the relative influences of gene flow and drift on the distribution of genetic variability. Evolution, 53: 1898-1914. Irion, G.; Kalliola, R. 2010. Long-term landscape development processes in Amazonia. In: Hoorn, C.; Wesselingh, F.P. (Eds). Amazonia, landscape and species evolution. Blackwell Publishing, Oxford. p. 185-197. 117 Kaefer, I.L.; Lima, A.P. 2012. Sexual signals of the Amazonian frog Allobates paleovarzensis: geographic variation and stereotypy of acoustic traits. Behaviour, 149: 15-33. Kaefer, I.L.; Montanarin, A.; Costa, R.S.; Lima, A.P. Temporal patterns of reproductive activity and site attachment of the Brilliant-thighed Frog Allobates femoralis from Central Amazonia. Journal of Herpetology, in press. Kaefer, I.L.; Tsuji-Nishikido, B.M.; Lima, A.P. 2012. Beyond the river: underlying determinants of population acoustic signal variability in Amazonian direct-developing Allobates (Anura: Dendrobatoidea). Acta Ethologica, doi:10.1007/s10211-012-0126-0. Kimura, M. 1980. A simple method for estimating evolutionary rate of base substitutions through comparative studies of nucleotide sequences. Journal of Molecular Evolution, 16: 111-120. Lemmon, E.M. 2009. Diversification of conspecific signals in sympatry: geographic overlap drives multidimensional reproductive character displacement in frogs. Evolution, 63: 11551170. Librado, P.; Rozas, J. 2009. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics, 25: 1451-1452. Lima, A.P.; Caldwell, J.P.; Biavati, G.; Montanarin, A. 2010. A new species of Allobates (Anura: Aromobatidae) from paleovárzea forest in Amazonas, Brazil. Zootaxa, 2337: 1-17. Losos, J.B. 2008. Phylogenetic niche conservatism, phylogenetic signal and the relationship between phylogenetic relatedness and ecological similarity among species. Ecology Letters, 11: 995-1003. Lougheed, S.C.; Austin, J.D.; Bogart, J.P.; Boag, P.T.; Chek, A.A. 2006. Multi-character perspectives on the evolution of intraspecific differentiation in a Neotropical hylid frog. BMC Evolutionary Biology, 6: 23. Lougheed, S.C.; Gascon, C.; Jones, D.A.; Bogart, J.P.; Boag, P.T. 1999. Ridges and rivers: a test of competing hypothesis of Amazonian diversification using a dart-poison frog (Epipedobates femoralis). Proceedings of the Royal Society B, 266: 1829-1835. Luna, A.G.; Hödl, W.; Amézquita, A. 2010. Colour, size and movement as visual subcomponents in multimodal communication by the frog Allobates femoralis. Animal Behaviour, 79: 739745. 118 Lynch, M.; Crease, T.J. 1990. The analysis of population survey data on DNA sequence variation. Molecular Biology and Evolution, 7: 377-394. Maan, M.E.; Cummings, M.E. 2009. Sexual dimorphism and directional sexual selection on aposematic signals in a poison frog. Proceedings of the National Academy of Sciences, 106: 19072-19077. Mantel, N. 1967. The detection of disease clustering and a generalized regression approach. Cancer Research, 27: 209-220. McClelland, B.E.; Wilczynski, W.; Ryan, M.J. 1996. Correlations between call characteristics and morphology in male cricket frogs (Acris crepitans). Journal of Experimental Biology, 199: 1907-1919. McKay, B.D.; Zink, R.M. 2009. The causes of mitochondrial DNA gene tree paraphyly in birds. Molecuar Phylogenetics and Evolution, 54: 647-650. Melendez, K.V.; Feng, A.S. 2010. Communication calls of little brown bats display individualspecific characteristics. The Journal of the Acoustic Society of America, 128: 919-923. Montanarin, A.; Kaefer, I.L.; Lima, A.P. 2011. Courtship and mating behaviour of the Brilliantthighed Frog Allobates femoralis from Central Amazonia: implications for the study of a species complex. Ethology Ecology & Evolution, 23: 141-150. Morales, V.R. 2002 “2000”. Sistemática y biogeografia del grupo trilineatus (Amphibia, Anura, Dendrobatidae, Colostethus), con descripción de once nuevas especies. Publicaciones de la Asociación Amigos de Doñana, 13: 1-59. Moritz, C.; Patton, J.L.; Schneider, C.J.; Smith, T.B. 2000. Diversification of rainforest faunas: an integrated molecular approach. Annual Review of Ecology and Systematics, 31: 533-563. Narins, P.M.; Grabul, D.S.; Soma, K.K.; Gaucher, P.; Hödl, W. 2005. Cross-modal integration in a dart-poison frog. Proceedings of the National Academy of Sciences, 102: 2425-2429. Narins, P.M.; Hödl, W.; Grabul, D.S. 2003. Bimodal signal requisite for agonistic behavior in a dart-poison frog, Epipedobates femoralis. Proceedings of the National Academy of Sciences, 100: 577-580. 119 Nielsen, R.; Beaumont, M.A. 2009. Statistical inferences in phylogeography. Molecular Ecology, 18: 1034-1047. Nosil, P.; Harmon, L.J.; Seehausen, O. 2009. Ecological explanations for (incomplete) speciation. Trends in Ecology and Evolution, 24: 145-156. Palumbi, S.R. 1996. Nucleic acids II: the polymerase chain reaction. In: Hillis, D.M.; Moritz, C.; Mable, B.K. (Eds). Molecular Systematics. Sinauer & Associates Inc., Sunderland, Massachusetts. p. 205-247. Panhuis, T.M.; Butlin, R.; Zuk, M.; Tregenza, T. 2001. Sexual selection and speciation. Trends in Ecology and Evolution, 16: 364-371. Patton, J.L.; Silva, M.N.F. 1998. Rivers, refuges, and ridges: The geography of speciation of Amazonian mammals. In: Howard, D.J; Berlocher, S.H. (Eds). Endless Forms: Species and Speciation. Oxford University Press, Oxford. p. 202-213. Patton, J.L.; Silva, M.N.F.; Malcolm, J.R. 1994. Gene genealogy and differentiation among arboreal spiny rats (Rodentia: Echymidae) of the Amazon Basin: a test of the riverine barrier hypothesis. Evolution, 48: 1314-1323. Peres, C.A.; Patton, J.L.; Silva, M.N.F. 1996. Riverine barriers and gene flow in Amazonian saddle-back tamarins. Folia Primatologica 67: 113-124. Prestwich, K.N. 1994. The energetics of acoustic signaling in anurans and insects. American Zoologist, 34: 625-643. Pröhl, H. 2005. Territorial behavior in dendrobatid frogs. Journal of Herpetology, 39: 354-365. Pröhl, H.; Hagemann, S.; Karsch, J.; Höbel, G. 2007. Geographic variation in male sexual signals in Strawberry Poison Frogs (Dendrobates pumilio). Ethology, 113: 825-837. Pröhl, H.; Koshy, R.A.; Mueller, U.; Rand, A.S.; Ryan, M.J. 2006. Geographic variation of genetic and behavioral traits in northern and southern túngara frogs. Evolution, 60: 16691679. Rambaut, A.; Drummond, A.J. 2007. Tracer v 1.5 (http://beast.bio.ed.ac.uk/Tracer). Acesso: 02/03/2012. 120 Ramos-Onsins, S.E.; Rozas, J. 2002. Statistical properties of new neutrality tests against population growth. Molecular Biology and Evolution, 19: 2092-2100. Ribas, C.C.; Aleixo, A.; Nogueira, A.C.R.; Miyaki, C.Y.; Cracraft, J. 2011. A palaeobiogeographic model for biotic diversification within Amazonia over the past three million years. Proceedings of the Royal Society B, 279: 681-689. Riddle, B.R. 2009. What is modern biogeography without phylogeography? Journal of Biogeography, 36: 1-2. Riddle, B.R.; Hafner, D.J. 2004. The past and future roles of phylogeography in historical biogeography. In: Lomolino, M.V.; Heaney, L.R. (Eds). Frontiers of biogeography: new directions in the geography of nature. Sinauer Associates Inc, Sunderland. p. 93-110. Rodríguez, A.; Nuez, D.; Alonso, R. 2010. Intraspecific variation in the advertisement call of the Cloud-Forest Frog Eleutherodactylus glamyrus (Anura: Eleutherodactylidae). Journal of Herpetology, 44: 457-466. Rosen, D.E. 1978. Vicariant patterns and historical explanation in biogeography. Systematic Zoology, 27: 159-188. Rossetti, D.F.; de Toledo, P.M.; Góes, A.M. 2005. New geological framework for Western Amazonia (Brazil) and implications for biogeography and evolution. Quaternary Research, 63: 78-89. Ryan, M.J. 1988. Energy, calling, and selection. American Zoologist, 28: 885-898. Ryan, M.J.; Cocroft, R.B.; Wilczynski, W. 1990. The role of environmental selection in intraspecific divergence of mate recognition signals in the cricket frog, Acris crepitans. Evolution, 44: 1869-1872. Ryan, M.J.; Rand, A.S.; Weigt, L.A. 1996. Allozyme and advertisement call variation in the túngara frog, Physalaemus pustulosus. Evolution, 50: 2435-2453. Ryan, M.J.; Wilczynski, W. 1991. Evolution of intraspecific variation in the advertisement call of a cricket frog (Acris crepitans, Hylidae). Biological Journal of the Linnean Society, 44: 249271. 121 Santini, F.; Miglietta, M.P.; Faucci, A. 2012. Where are we now? An introduction to a special issue on speciation. Evolutionary Biology, 39: 141-147. Santos, J.C.; Coloma, L.A.; Summers, K.; Caldwell, J.P.; Ree, R.; Cannatella, D.C. 2009. Amazonian amphibian diversity is primarily derived from Late Miocene Andean lineages. PLOS Biology, 7: e1000056. Santos, S.; Hrbek, T.; Farias, I.P.; Schneider, H.; Sampaio, I. 2006. Population genetic structuring of the king weakfish, Macrodon ancylodon (Sciaenidae), in Atlantic coastal waters of South America: deep genetic divergence without morphological change. Molecular Ecology, 15: 4361-4373. Schneider, S.; Excoffier, L. 1999. Estimation of demographic parameters from the distribution of pairwise differences when the mutation rates vary among sites: Application to human mitochondrial DNA. Genetics, 152: 1079-1089. Sequeira, F.; Sodré, D.; Ferrand, N.; Bernardi, J.; Sampaio, I.; Schneider, H.; Vallinoto, M. 2011. Hybridization and massive mtDNA unidirectional introgression between the closely related Neotropical toads Rhinella marina and R. schneideri inferred from mtDNA and nuclear markers. BMC Evolutionary Biology, 11: 264. Silva, M.N.F.; Patton, J.L. 1998. Molecular phylogeography and the evolution and conservation of mammals. Molecular Ecology, 7: 475-486. Simões, P.I. 2010. Diversificação do complexo Allobates femoralis (Anura, Dendrobatidae) em florestas da Amazônia brasileira: desvendando padrões atuais e históricos. Tese de Doutorado, Instituto Nacional de Pesquisas da Amazônia, Manaus, Amazonas. Simões, P.I.; Lima, A.P.; Magnusson, W.E.; Hödl, W.; Amézquita, A. 2008. Acoustic and morphological differentiation in the frog Allobates femoralis: Relationships with the upper Madeira River and other potential geological barriers. Biotropica, 40: 607-614. Sioli, H. 1984. The Amazon: Limnology and Landscape Ecology of a Mighty Tropical River and its Basin. Dr W. Junk Publisher, Dordrecht. Smith, M.J.; Hunter, D. 2005. Temporal and geographic variation in the advertisement call of the booroolong frog (Litoria booroolongensis: Anura: Hylidae). Ethology, 111: 1103-1115. 122 Smouse, P.E.; Long, J.C.; Sokal, R.R. 1986. Multiple regression and correlation extensions of the Mantel test of matrix correspondence. Systematic Zoology, 35: 727-732. Solomon, S.E.; Bacci, M.; Martins, J.; Vinha, G.G.; Mueller, U.G. 2008. Paleodistributions and comparative molecular phylogeography of leafcutter ants (Atta spp.) provide new insight into the origins of Amazonian diversity. PLoS One, 3: e2738. Tajima, F. 1989. Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics, 123: 585-595. Tamura, K.; Peterson, D.; Peterson, N.; Stecher, G.; Nei, M.; Kumar, S. 2011. MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Molecular Biology and Evolution, 28: 2731-2739. Tárano, Z. 2001. Variation in male advertisement calls in the neotropical frog Physalaemus enesefae. Copeia, 2001: 1064-1072. Telles, M.P.C.; Silva, R.S.M.; Chaves, L.J.; Coelho, A.S.G.; Diniz–Filho, J.A.F. 2001. Divergência entre subpopulações de cagaiteira (Eugenia dysenterica) em resposta a padrões edáficos e distribuição espacial. Pesquisa Agropecuária Brasileira, 36: 1387-1394. Templeton, A.R.; Crandall, K.A.; Sing, C.F. 1992. A cladistic analysis of phenotypic associations with haplotypes inferred from restriction endonuclease mapping and DNA sequence data. III. Cladogram estimation. Genetics, 132: 619-633. Thompson, J.D.; Higgins, D.G.; Gibson, T.J. 1994. Improved sensitivity of profile searches through the use of sequence weights and gap excision. Computer Applications in the Biosciences, 10: 19-29. Tsuji-Nishikido, B.; Kaefer, I.L.; Freitas, F.C.; Menin, M.; Lima, A.P. 2012. Significant but not diagnostic: Differentiation through morphology and calls in the Amazonian frogs Allobates nidicola and A. masniger. Herpetological Journal, 22: 105-114. Turmelle, A.S.; Kunz, T.H.; Sorenson, M.D. 2011. A tale of two genomes: contrasting patterns of phylogeographic structure in a widely distributed bat. Molecular Ecology, 20: 357-375. Vences, M.; Wake, D.B. 2007. Speciation, species boundaries and phylogeography of amphibians. In: Heatwole, H.; Tyler, M.J. (Eds). Amphibian Biology, Systematics. Surrey Beatty & Sons, Chipping Norton. p. 2613-2671. 123 Vences, M.; Kosuch, J.; Lötters, S.; Widmer, A.; Jungfer, K.; Köhler, J.; Veith, M. 2000. Phylogeny and classification of Poison Frogs (Amphibia: Dendrobatidae), based on mitochondrial 16S and 12S ribosomal RNA gene sequences. Molecular Phylogenetics and Evolution, 15: 34-40. Vences, M.; Thomas, M.; Meijden, A.; Chiari, Y.; Vieites, D.R. 2005. Comparative performance of the 16S rRNA gene in DNA barcoding of amphibians. Frontiers in Zoology, 2: 5. Verburgt, L.; Ferreira, M.; Ferguson, J.W.H. 2011. Male field cricket song reflects age, allowing females to prefer young males. Animal Behaviour, 81: 19-29. Verdade, V.K.; Rodrigues, M.T. 2007. Taxonomic review of Allobates (Anura, Aromobatidae) from the Atlantic Forest, Brazil. Journal of Herpetology, 41: 566-580. Vieites, D.R.; Wollenberg, K.C.; Adreone, F.; Köhler, J.; Glaw, F.; Vences, M. 2009. Vast underestimation of Madagascar’s biodiversity evidenced by an integrative amphibian inventory. Proceedings of the National Academy of Sciences, 106: 8267-8272. Wallace, A.R. 1852. On the monkeys of the Amazon. Proceedings of the Zoological Society of London, 20: 107-110. Wells, K.D. 2007. The ecology and behavior of amphibians. The University of Chicago Press, Chicago. Wells, K.D.; Schwartz, J.J. 2007. The Behavioral Ecology of Anuran Communication. In: Narins, P.M.; Feng, A.S.; Fay, R.R.; Popper, A.N. (Eds). Hearing and Sound Communication in Amphibians. Springer-Verlag, New York. p. 44-86. West-Eberhard, M.J. 1989. Phenotypic plasticity and the origins of diversity. Annual Review of Ecology and Systematics, 20: 249-278. Wiens, J.J. 2004. Speciation and ecology revisited: phylogenetic niche conservatism and the origin of species. Evolution, 58: 193-197. Wiens, J.J. 2007. Species delimitation: New approaches for discovering diversity. Systematic Biology, 56: 875-878. Wiens, J.J. 2008. Systematics and herpetology in the age of genomics. Bioscience, 58: 297-307. Wiley, E.O. 1978. The evolutionary species concept reconsidered. Systematic Zoology, 27: 17-26. 124 Wilkinson, L. 2007. SYSTAT: the system for statistics. SYSTAT inc., Evanston. Wright, S. 1951. The genetical structure of populations. Annals of Human Genetics, 15: 323-354. Wycherley, J.; Doran, S.; Beebee, T.J.C. 2002. Male advertisement call characters as phylogeographical indicators in European water frogs. Biological Journal of the Linnean Society, 77: 355-365. Zeisset, I.; Beebee, T.J.C. 2008. Amphibian phylogeography: a model for understanding historical aspects of species distributions. Heredity, 101: 109-119. Zhang, D.X.; Hewitt, G.M. 2003. Nuclear DNA analyses in genetic studies of populations: practice, problems and prospects. Molecular Ecology, 12: 563-584. Zink, R.M.; Barrowclough, G.F. 2008. Mitochondrial DNA under siege in avian phylogeography. Molecular Ecology, 17: 2107-2121. Zink, R.M.; Blackwell-Rago, R.C.; Ronquist, F. 2000. The shifting roles of vicariance in biogeography. Proceedings of the Royal Society B, 267: 497-503. Zweifel, R.G. 1968. Effects of temperature, body size, and hybridization on mating calls of toads, Bufo a. americanus and Bufo woodhousii fowleri. Copeia, 1968: 269-285. 125 Anexos Pareceres emitidos pelas bancas examinadoras da aula de qualificação, da versão escrita da tese e da defesa pública da tese, respectivamente. 126 127 128 129 130 131