Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-31T13:09:41.367Z Has data issue: false hasContentIssue false

Rapid divergence, molecular evolution, and morphological diversification of coastal host-parasite systems from southern Brazil

Published online by Cambridge University Press:  20 June 2019

Marlus Bueno-Silva*
Affiliation:
Laboratório de Ecologia Molecular e Parasitologia Evolutiva (LEMPE), Departamento de Zoologia, Centro Politécnico, Setor de Ciências Biológicas, Universidade Federal do Paraná (UFPR), Caixa Postal 19073, CEP 81531-980, Curitiba, Paraná, Brazil Programa de Pós-Graduação em Zoologia, Departamento de Zoologia, Setor de Ciências Biológicas, Universidade Federal do Paraná (UFPR), Caixa Postal 19020, CEP 81531-980, Curitiba, Paraná, Brazil
Walter A. Boeger
Affiliation:
Laboratório de Ecologia Molecular e Parasitologia Evolutiva (LEMPE), Departamento de Zoologia, Centro Politécnico, Setor de Ciências Biológicas, Universidade Federal do Paraná (UFPR), Caixa Postal 19073, CEP 81531-980, Curitiba, Paraná, Brazil
*
Author for correspondence: Marlus Bueno-Silva, E-mail: marbsilva@gmail.com

Abstract

This study assessed the role of historical processes on the geographic isolation, molecular evolution, and morphological diversification of host-parasite populations from the southern Brazilian coast. Adult specimens of Scleromystax barbatus and Scleromystax macropterus were collected from the sub-basin of the Nhundiaquara River and the sub-basin of the Paranaguá Bay, state of Paraná, Brazil. Four species of Gyrodactylus were recovered from the body surface of both host species. Morphometric analysis of Gyrodactylus spp. and Scleromystax spp. indicated that subpopulations of parasites and hosts could be distinguished from different sub-basins and locations, but the degree of morphological differentiation seems to be little related to geographic distance between subpopulations. Phylogenetic relationships based on DNA sequences of Gyrodactylus spp. and Scleromystax spp. allowed distinguishing lineages of parasites and hosts from different sub-basins. However, the level of genetic structuring of parasites was higher in comparison to host species. Evidence of positive selection in mtDNA sequences is likely associated with local adaptation of lineages of parasites and hosts. A historical demographic analysis revealed that populations of Gyrodactylus and Scleromystax have expanded in the last 250 000 years. The genetic variation of parasites and hosts is consistent with population-specific selection, population expansions, and recent evolutionary co-divergence.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2019 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Ab'Saber, AN (2000) Fundamentos da geomorfologia costeira do Brasil Atlântico inter e subtropical. Revista Brasileira de Geomorfologia 1, 2743.Google Scholar
Agrawal, AA (2001) Phenotypic plasticity in the interactions and evolution of species. Science 294, 321326.Google Scholar
Allender, CJ, Seehausen, O, Knight, ME, Turner, GF and Maclean, N (2003) Divergent selection during speciation of Lake Malawi cichlid fishes inferred from parallel radiations in nuptial coloration. Proceedings of the National Academy of Sciences of the United States of America 100, 1407414079.Google Scholar
Almeida, FFM (1976) The system of continental rifts bordering the Santos basin, Brazil. Anais da Academia Brasileira de Ciências 48, 1526.Google Scholar
Angulo, RJ (2004) Mapa do Cenozoico do litoral do estado do Paraná. Boletim Paranaense de Geociências 55, 2542.Google Scholar
Angulo, RJ and Lessa, GC (1997) The Brazilian sea-level curves: a critical review with emphasis on the curves from the Paranaguá and Cananéia regions. Marine Geology 140, 141166.Google Scholar
Angulo, RJ, Pessenda, LC and Souza, MC (2002) O significado das datações ao 14C na reconstrução de paleoníveis marinhos e na evolução das barreiras quaternárias do litoral paranaense. Revista Brasileira de Geociências 32, 95106.Google Scholar
Araujo, SBL, Braga, MP, Brooks, DR, Agosta, SJ, Hoberg, EP, von Hartenthal, FW and Boeger, WA (2015) Understanding host-switching by ecological fitting. PLoS ONE 10, 117. http://dx.doi.org/10.1371/journal.pone.0139225.Google Scholar
Árnason, E and Rand, DM (1992) Heteroplasmy of short tandem repeats in mitochondrial DNA of Atlantic cod, Gadus morhua. Genetics 132, 211220.Google Scholar
Avise, JC (2000) Phylogeography: The History and Formation of species. Massachusetts, USA: Harvard University Press.Google Scholar
Bakke, TA, Harris, PD and Cable, J (2002) Host specificity dynamics: observations on gyrodactylid monogeneans. International Journal for Parasitology 32, 281308.Google Scholar
Bakke, TA, Cable, J and Harris, PD (2007) The biology of gyrodactylid monogeneans: the ‘Russian doll-killers’. Advances in Parasitology 64, 161376.Google Scholar
Bankers, L, Fields, P, McElroy, KE, Boore, JL, Logsdon, JM Jr and Neiman, M (2017) Genomic evidence for population-specific responses to co-evolving parasites in a New Zealand freshwater snail. Molecular Ecology 26, 36633675.Google Scholar
Bazin, E, Glémin, S and Galtier, N (2006) Population size does not influence mitochondrial genetic diversity in animals. Science 312, 570572.Google Scholar
Beheregaray, LB, Sunnucks, P and Briscoe, DA (2002) A rapid fish radiation associated with the last sea-level changes in southern Brazil: the silverside Odontesthes perugiae complex. Proceedings of the Royal Society of London B 269, 6573.Google Scholar
Benson, G (1999) Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Research 27, 573580.Google Scholar
Bigarella, JJ (2001) Contribuição ao estudo da planície litorânea do estado do Paraná. Brazilian Archives of Biology and Technology Jubilee, 65110.Google Scholar
Boeger, WA, Kritsky, DC and Pie, MR (2003) Context of diversification of the viviparous Gyrodactylidae (Platyhelminthes, Monogenoidea). Zoologica Scripta 32, 437448.Google Scholar
Boeger, WA, Kritsky, DC, Pie, MR and Engers, KB (2005) Mode of transmission, host switching, and escape from the Red Queen by viviparous gyrodactylids (Monogenoidea). Journal of Parasitology 91, 10001007.Google Scholar
Boeger, WA, Vianna, RT and Thatcher, VE (2006) Monogenoidea. In Adis, J, Arias, JR, Rueda-Delgado, G and Wantzen, KM (eds). Aquatic Biodiversity in Latin America, 2nd Edn. vol. 1, Amazon Fish Parasites, Sofia, Bulgaria: Pensoft Publishers, pp. 42116.Google Scholar
Bonnet, E and Van de Peer, Y (2002) ZT: a software tool for simple and partial Mantel tests. Journal of Statistical Software 7, 112.Google Scholar
Bouckaert, R, Heled, J, Kühnert, D, Vaughan, T, Wu, C-H, Xie, D, Suchard, MA, Rambaut, A and Drummond, AJ (2014) BEAST 2: a software platform for Bayesian evolutionary analysis. PLoS Computational Biology 10, 16. http://dx.doi.org/10.1371/journal.pcbi.1003537.Google Scholar
Brown, WM, George, M Jr and Wilson, AC (1979) Rapid evolution of animal mitochondrial DNA. Proceedings of the National Academy of Sciences of the United States of America 76, 19671971.Google Scholar
Bueno-Silva, M and Boeger, WA (2009) Neotropical Monogenoidea. 53. Gyrodactylus corydori sp. n. and redescription of Gyrodactylus anisopharynx (Gyrodactylidea: Gyrodactylidae), parasites of Corydoras spp. (Siluriformes: Callichthyidae) from southern Brazil. Folia Parasitologica 56, 1320.Google Scholar
Bueno-Silva, M and Boeger, WA (2014) Neotropical Monogenoidea. 58. Three new species of Gyrodactylus (gyrodactylidae) from Scleromystax spp. (Callichthyidae) and the proposal of COII gene as an additional fragment for barcoding gyrodactylids. Folia Parasitologica 61, 213222.Google Scholar
Bueno-Silva, M, Boeger, WA and Pie, MR (2011) Choice matters: incipient speciation in Gyrodactylus corydori (Monogenoidea: Gyrodactylidae). International Journal for Parasitology 41, 657667.Google Scholar
Bush, AO, Lafferty, KD, Lotz, JM, Shostak, AW (1997) Parasitology meets ecology on its own terms: Margolis et al. revisited. Journal of Parasitology 83, 575583.Google Scholar
Cable, J and Harris, PD (2002) Gyrodactylid developmental biology: historical review, current status and future trends. International Journal for Parasitology 32, 255280.Google Scholar
Caldera, EJ and Bolnick, DI (2008) Effects of colonization history and landscape structure on genetic variation within and among threespine stickleback (Gasterosteus aculeatus) populations in a single watershed. Evolutionary Ecology Research 10, 575598.Google Scholar
Castellana, S, Vicario, S and Saccone, C (2011) Evolutionary patterns of the mitochondrial genome in Metazoa: exploring the role of mutation and selection in mitochondrial protein-coding genes. Genome Biology and Evolution 3, 10671079.Google Scholar
Chaklader, MR, Siddik, MAB, Hanif, MA, Nahar, A, Mahmud, S and Piria, M (2016) Morphometric and meristic variation of endangered pabda catfish, Ompok pabda (Hamilton-Buchanan, 1822) from southern coastal waters of Bangladesh. Pakistan Journal of Zoology 48, 681687.Google Scholar
Chen, CA, Ablan, MCA, McManus, JW, Bell, JD, Tuan, VS, Cabanban, AS and Shao, K-T (2004) Variable numbers of tandem repeats (VNTRs), heteroplasmy, and sequence variation of the mitochondrial control region in the threespot Dascyllus, Dascyllus trimaculatus (Perciformes: Pomacentridae). Zoological Studies 43, 803812.Google Scholar
Cohen, SC and Kohn, A (2008) South American monogenea – list of species, hosts and geographical distribution from 1997 to 2008. Zootaxa 1924, 142.Google Scholar
Comesaña, AS, Martínez-Areal, MT and Sanjuan, A (2008) Genetic variation in the mitochondrial DNA control region among horse mackerel (Trachurus trachurus) from the Atlantic and Mediterranean areas. Fisheries Research 89, 122131.Google Scholar
Consuegra, S, John, E, Verspoor, E and Leaniz, CG (2015) Patterns of natural selection acting on the mitochondrial genome of a locally adapted fish species. Genetics, Selection, Evolution 47, 110. http://dx.doi.org/10.1186/s12711-015-0138-0.Google Scholar
Crispo, E, Bentzen, P, Reznick, DN, Kinnison, MT and Hendry, AP (2006) The relative influence of natural selection and geography on gene flow in guppies. Molecular Ecology 15, 4962.Google Scholar
Crookes, S and Shaw, PW (2016) Isolation by distance and non-identical patterns of gene flow within two river populations of the freshwater fish Rutilus rutilus (L. 1758). Conservation Genetics 17, 861874.Google Scholar
Darriba, D, Taboada, GL, Doallo, R and Posada, D (2012) Jmodeltest 2: more models, new heuristics and parallel computing. Nature Methods 9, 772.Google Scholar
Dávidová, M, Jarkovský, J, Matejusová, I and Gelnar, M (2005) Seasonal occurrence and metrical variability of Gyrodactylus rhodei Žitnan 1964 (Monogenea, Gyrodactylidae). Parasitology Research 95, 398405.Google Scholar
Dawson, AG (1992) Ice Age Earth: Late Quaternary geology and climate. London, UK: Routledge.Google Scholar
Desdevises, Y, Morand, S, Jousson, O and Legendre, P (2002) Coevolution between Lamellodiscus (monogenea: Diplectanidae) and Sparidae (Teleostei): the study of a complex host-parasite system. Evolution 56, 24592471.Google Scholar
Dmitrieva, E and Dimitrov, G (2002) Variability in the taxonomic characters of Black Sea gyrodactylids (Monogenea). Systematic Parasitology 51, 199206.Google Scholar
Drummond, AJ, Ho, SYW, Phillips, MJ and Rambaut, A (2006) Relaxed phylogenetics and dating with confidence. PLoS Biology 4, 699710.Google Scholar
Du Preez, LH and Maritz, MF (2006) Demonstrating morphometric protocols using polystome marginal hooklet measurements. Systematic Parasitology 63, 115.Google Scholar
Dybdahl, MF and Storfer, A (2003) Parasite local adaptation: Red Queen versus Suicide King. Trends in Ecology & Evolution 18, 523530.Google Scholar
Edgar, RC (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research 32, 17921797.Google Scholar
Excoffier, L and Lischer, HEL (2010) Arlequin suite ver 3.5: a new series of programs to perform population genetics analyses under Linux and Windows. Molecular Ecology Resources 10, 564567.Google Scholar
Faber, JE and Stepien, CA (1998) Tandemly repeated sequences in the mitochondrial DNA control region and phylogeography of the Pike-Perches Stizostedion. Molecular Phylogenetics and Evolution 10, 310322.Google Scholar
Fay, JC and Wu, C-I (2000) Hitchhiking under positive Darwinian selection. Genetics 155, 14051413.Google Scholar
Ferraris, CJ (2007) Checklist of catfishes, recent and fossil (Osteichthyes: Siluriformes), and catalogue of siluriform primary types. Zootaxa 1418, 1628.Google Scholar
Fischer, C, Koblmüller, S, Gülly, C, Schlötterer, C, Sturmbauer, C and Thallinger, GG (2013) Complete mitochondrial DNA sequences of the threadfin cichlid (Petrochromis trewavasae) and the blunthead cichlid (Tropheus moorii) and patterns of mitochondrial genome evolution in cichlid fishes. PLoS ONE 8, 114. http://dx.doi.org/10.1371/journal.pone.0067048.Google Scholar
Gagnon, M-C and Angers, B (2006) The determinant role of temporary proglacial drainages on the genetic structure of fishes. Molecular Ecology 15, 10511065.Google Scholar
Gandon, S and Michalakis, Y (2002) Local adaptation, evolutionary potential and host-parasite coevolution: interactions between migration, mutation, population size and generation time. Journal of Evolutionary Biology 15, 451462.Google Scholar
Garvin, MR, Bielawski, JP and Gharrett, AJ (2011) Positive Darwinian selection in the piston that powers proton pumps in complex I of the mitochondria of Pacific salmon. PLoS ONE 6, 115, http://dx.doi.org/10.1371/journal.pone.0024127.Google Scholar
Geets, A, Appleby, C and Ollevier, F (1999) Host-dependent and seasonal variation in opisthaptoral hard parts of Gyrodactylus cf. arcuatus from three Pomatoschistus spp. and G. arcuatus from Gasterosteus aculeatus: a multivariate approach. Parasitology 119, 2740.Google Scholar
Glennon, V, Perkins, EM, Chisholm, LA and Whittington, ID (2008) Comparative phylogeography reveals host generalists, specialists and cryptic diversity: hexabothriid, microbothriid and monocotylid monogeneans from rhinobatid rays in southern Australia. International Journal for Parasitology 38, 15991612.Google Scholar
Guimarães, ATB, Menezes, MS and Peret, AC (2010) Composição da ictiofauna em função da fisiografia de um riacho costeiro de Floresta Atlântica – Brasil. Biota Neotropica 10, 5765.Google Scholar
Gunawickrama, KBS (2007) Morphological heterogeneity in some estuarine populations of the catfish Arius jella (Ariidae) in Sri Lanka. Ceylon Journal of Science (Biological Sciences) 36, 100107.Google Scholar
Hänfling, B, Durka, W and Brandl, R (2004) Impact of habitat fragmentation on genetic population structure of roach, Rutilus rutilus, in a riparian ecosystem. Conservation Genetics 5, 247257.Google Scholar
Hardouin, EA and Tautz, D (2013) Increased mitochondrial mutation frequency after an island colonization: positive selection or accumulation of slightly deleterious mutations? Biology Letters 9, 15. http://dx.doi.org/10.1098/rsbl.2012.1123.Google Scholar
Harris, PD (1998) Extreme morphological variation between related individuals of Gyrodactylus pungitti Malmberg, 1964 (Monogenea). Systematic Parasitology 39, 137140.Google Scholar
Harris, PD, Shinn, AP, Cable, J, Bakke, TA and Bron, JE (2008) Gyrodb: gyrodactylid monogeneans on the web. Trends in Parasitology 24, 109111.Google Scholar
Hayakawa, T, Culleton, R, Otani, H, Horii, T and Tanabe, K (2008) Big Bang in the evolution of extant malaria parasites. Molecular Biology and Evolution 25, 22332239.Google Scholar
Hewitt, G (2000) The genetic legacy of the quaternary ice ages. Nature 405, 907913.Google Scholar
Hickerson, MJ, Carstens, BC, Cavender-Bares, J, Crandall, KA, Graham, CH, Johnson, JB, Rissler, L, Victoriano, PF and Yoder, AD (2010) Phylogeography's past, present, and future: 10 years after Avise, 2000. Molecular Phylogenetics and Evolution 54, 291301.Google Scholar
Hoarau, G, Holla, S, Lescasse, R, Stam, WT and Olsen, JL (2002) Heteroplasmy and evidence for recombination in the mitochondrial control region of the flatfish Platichthys flesus. Molecular Biology and Evolution 19, 22612264.Google Scholar
Hoberg, EP and Brooks, DR (2015) Evolution in action: climate change, biodiversity dynamics and emerging infectious disease. Philosophical Transactions of the Royal Society B 370, 17. http://dx.doi.org/10.1098/rstb.2013.0553.Google Scholar
Horodesky, A, Abilhoa, V, Zeni, TO, Neto, RM, Castilho-Westphal, GG and Ostrensky, A (2015) Ecological analysis of the ichthyofaunal community ten years after a diesel oil spill at Serra do Mar, Paraná state, Brazil. Global Ecology and Conservation 4, 311320.Google Scholar
Huyse, T and Volckaert, FAM (2005) Comparing host and parasite phylogenies: Gyrodactylus flatworms jumping from Goby to Goby. Systematic Biology 54, 710718.Google Scholar
Huyse, T, Audenaert, V and Volckaert, FAM (2003) Speciation and host-parasite relationships in the genus Gyrodactylus (monogenea, Platyhelminthes) infecting gobies of the genus Pomatoschistus (gobiidae, Teleostei). International Journal for Parasitology 33, 16791689.Google Scholar
Imbrie, J, Boyle, EA, Clemens, SC, Duffy, A, Howard, WR, Kukla, G, Kutzbach, J, Martinson, DG, McIntyre, A, Mix, AC, Molfino, B, Morley, JJ, Peterson, LC, Pisias, NG, Prell, WL, Raytoo, ME, Shackletons, NJ and Toggweile, JR (1992) On the structure and origin of major glaciation cycles. 1. Linear responses to Milankovitch forcing. Paleoceanography 7, 701738.Google Scholar
Imre, I, McLaughlin, RL and Noakes, DLG (2002) Phenotypic plasticity in brook charr: changes in caudal fin induced by water flow. Journal of Fish Biology 61, 11711181.Google Scholar
Jamandre, BW, Durand, J-D and Tzeng, W-N (2014) High sequence variations in mitochondrial DNA control region among worldwide populations of flathead mullet Mugil cephalus. International Journal of Zoology 2014, 110. http://dx.doi.org/10.1155/2014/564105.Google Scholar
James, JE, Piganeau, G and Eyre-Walker, A (2016) The rate of adaptive evolution in animal mitochondria. Molecular Ecology 25, 6778.Google Scholar
Jarman, SN, Ward, RD and Elliott, NG (2002) Oligonucleotide primers for PCR amplification of coelomate introns. Marine Biotechnology 4, 347355.Google Scholar
Kaatz, IM and Lobel, PS (1999) Acoustic behavior and reproduction in five species of Corydoras catfishes (Callichthyidae). The Biological Bulletin 197, 241242.Google Scholar
Kaltz, O and Shykoff, JA (1998) Local adaptation in host-parasite systems. Heredity 81, 361370.Google Scholar
Kita, K, Hirawake, H and Takamiya, S (1997) Cytochromes in the respiratory chain of helminth mitochondria. International Journal for Parasitology 27, 617630.Google Scholar
Kohn, A and Cohen, SC (1998) South American monogenea – list of species, hosts and geographical distribution. International Journal for Parasitology 28, 15171554.Google Scholar
Kotlík, P and Berrebi, P (2001) Phylogeography of the barbel (Barbus barbus) assessed by mitochondrial DNA variation. Molecular Ecology 10, 21772185.Google Scholar
Langerhans, RB, Layman, CA, Langerhans, AK and Dewitt, TJ (2003) Habitat-associated morphological divergence in two Neotropical fish species. Biological Journal of the Linnean Society 80, 689698.Google Scholar
Lecomte, F, Grant, WS, Dodson, JJ, Rodríguez-Sánchez, R and Bowen, BW (2004) Living with uncertainty: genetic imprints of climate shifts in East Pacific anchovy (Engraulis mordax) and sardine (Sardinops sagax). Molecular Ecology 13, 21692182.Google Scholar
Lee, W-J, Conroy, J, Howell, WH and Kocher, TD (1995) Structure and evolution of teleost mitochondrial control regions. Journal of Molecular Evolution 41, 5466.Google Scholar
Librado, P and Rozas, J (2009) DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics (Oxford, England) 25, 14511452.Google Scholar
Liu, J-X, Gao, T-X, Yokogawa, K and Zhang, Y-P (2006) Differential population structuring and demographic history of two closely related fish species, Japanese sea bass (Lateolabrax japonicus) and spotted sea bass (Lateolabrax maculatus) in Northwestern Pacific. Molecular Phylogenetics and Evolution 39, 799811.Google Scholar
Lleonart, J, Salat, J and Torres, GJ (2000) Removing allometric effects of body size in morphological analysis. Journal of Theoretical Biology 205, 8593.Google Scholar
Ludwig, A, May, B, Debus, L and Jenneckens, I (2000) Heteroplasmy in the mtDNA control region of sturgeon (Acipenser, Huso and Scaphirhynchus). Genetics 156, 19331947.Google Scholar
Maack, R (2001) Breves notícias sobre a geologia dos estados do Paraná e Santa Catarina. Brazilian Archives of Biology and Technology. Jubilee, 169288.Google Scholar
Maddison, WP and Maddison, DR (2010) Mesquite: a modular system for evolutionary analysis. Version 2.74. http://www.mesquiteproject.org (accessed 24 March 2011).Google Scholar
Malmberg, G (1970) The excretory systems and the marginal hooks as a basis for the systematic of Gyrodactylus (trematoda, Monogenea). Arkiv för Zoologi 23, 1235.Google Scholar
Mantel, N (1967) The detection of disease clustering and a generalized regression approach. Cancer Research 27, 209220.Google Scholar
Mantel, N and Valand, RS (1970) A technique of nonparametric multivariate analysis. Biometrics 26, 547558.Google Scholar
Martin, L, Suguio, K, Flexor, J-M, Dominguez, JML and Bittencourt, ACSP (1996) Quaternary sea-level history and variation in dynamics along the central Brazilian coast: consequences on coastal plain construction. Anais da Academia Brasileira de Ciências 68, 303354.Google Scholar
McMillan, WO and Palumbi, SR (1997) Rapid rate of control-region evolution in Pacific butterflyfishes (Chaetodontidae). Journal of Molecular Evolution 45, 473484.Google Scholar
McVicar, AH (1997) Disease and parasite implications of the coexistence of wild and cultured Atlantic salmon populations. ICES Journal of Marine Science 54, 10931103.Google Scholar
McWilliam, H, Li, W, Uludag, M, Squizzato, S, Park, YM, Buso, N, Cowley, AP and Lopez, R (2013) Analysis tool web services from the EMBL-EBI. Nucleic Acids Research 41, W597W600.Google Scholar
Meinilä, M, Kuusela, J, Ziętara, MS and Lumme, J (2004) Initial steps of speciation by geographic isolation and host switch in salmonid pathogen Gyrodactylus salaris (Monogenea: Gyrodactylidae). International Journal for Parasitology 34, 515526.Google Scholar
Melo-Ferreira, J, Vilela, J, Fonseca, MM, Fonseca, RR, Boursot, P and Alves, PC (2014) The elusive nature of adaptive mitochondrial DNA evolution of an Arctic lineage prone to frequent introgression. Genome Biology and Evolution 6, 886896.Google Scholar
Mo, TA (1991 a) Variations of opisthaptoral hard parts of Gyrodactylus salaris Malmberg, 1957 (Monogenea: Gyrodactylidae) on parr of Atlantic salmon Salmo salar L. in laboratory experiments. Systematic Parasitology 20, 1119.Google Scholar
Mo, TA (1991 b) Variations of opisthaptoral hard parts of Gyrodactylus salaris Malmberg, 1957 (Monogenea: Gyrodactylidae) on rainbow trout Oncorhynchus mykiss (Walbaum, 1792) in a fish farm, with comments on the spreading of the parasite in south-eastern Norway. Systematic Parasitology 20, 19.Google Scholar
Montoya-Burgos, JI (2003) Historical biogeography of the catfish genus Hypostomus (siluriformes: Loricariidae), with implications on the diversification of Neotropical ichthyofauna. Molecular Ecology 12, 18551867.Google Scholar
Morand, S, Simková, A, Matejusová, I, Plaisance, L, Verneau, O and Desdevises, Y (2002) Investigating patterns may reveal processes: evolutionary ecology of ectoparasitic monogeneans. International Journal for Parasitology 32, 111119.Google Scholar
Nesbø, CL, Fossheim, T, Vøllestad, LA and Jakobsen, KS (1999) Genetic divergence and phylogeographic relationships among European perch (Perca fluviatilis) populations reflect glacial refugia and postglacial colonization. Molecular Ecology 8, 13871404.Google Scholar
Nieberding, C, Morand, S, Libois, R and Michaux, JR (2004) A parasite reveals cryptic phylogeographic history of its host. Proceedings of the Royal Society of London B 271, 25592568.Google Scholar
Nieberding, CM, Durette-Desset, M-C, Vanderpoorten, A, Casanova, JC, Ribas, A, Deffontaine, V, Feliu, C, Morand, S, Libois, R and Michaux, JR (2008) Geography and host biogeography matter for understanding the phylogeography of a parasite. Molecular Phylogenetics and Evolution 47, 538554.Google Scholar
Nuismer, SL (2006) Parasite local adaptation in a geographic mosaic. Evolution 60, 2430.Google Scholar
Olstad, K, Bachmann, L and Bakke, TA (2009) Phenotypic plasticity of taxonomic and diagnostic structures in gyrodactylosis-causing flatworms (Monogenea, Platyhelminthes). Parasitology 136, 13051315.Google Scholar
Paxton, CGM (1997) Shoaling and activity levels in Corydoras. Journal of Fish Biology 51, 469502.Google Scholar
Pelletier, F, Garant, D and Hendry, AP (2009) Eco-evolutionary dynamics. Philosophical Transactions of the Royal Society B 364, 14831489.Google Scholar
Petit, JR, Jouzel, J, Raynaud, D, Barkov, NI, Barnola, J-M, Basile, I, Bender, M, Chappellaz, J, Davisk, M, Delaygue, G, Delmotte, M, Kotlyakov, VM, Legrand, M, Lipenkov, VY, Lorius, C, Pépin, L, Ritz, C, Saltzmank, E and Stievenard, M (1999) Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399, 429436.Google Scholar
Pettersen, RA, Mo, TA, Hansen, H and Vøllestad, LA (2015) Genetic population structure of Gyrodactylus thymalli (Monogenea) in a large Norwegian river system. Parasitology 142, 16931702.Google Scholar
Ponce, M, Infante, C, Jiménez-Cantizano, RM, Pérez, L and Manchado, M (2008) Complete mitochondrial genome of the blackspot seabream, Pagellus bogaraveo (Perciformes: Sparidae), with high levels of length heteroplasmy in the WANCY region. Gene 409, 4452.Google Scholar
Poulin, R and Morand, S (1999) Geographical distances and the similarity among parasite communities of conspecific host populations. Parasitology 119, 369374.Google Scholar
Prugnolle, F, Theron, A, Pointier, JP, Jabbour-Zahab, R, Jarne, P, Durand, P and de Meeûs, T (2005) Dispersal in a parasitic worm and its two hosts: consequence for local adaptation. Evolution 59, 296303.Google Scholar
Raeymaekers, JAM, Maes, GE, Audenaert, E and Volckaert, FAM (2005) Detecting Holocene divergence in the anadromous-freshwater three-spined stickleback (Gasterosteus aculeatus) system. Molecular Ecology 14, 10011014.Google Scholar
Rambaut, A, Drummond, AJ, Xie, D, Baele, G and Suchard, MA (2018) Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Systematic Biology 67, 901904.Google Scholar
Ramos-Onsins, SE and Rozas, J (2002) Statistical properties of new neutrality tests against population growth. Molecular Biology and Evolution 19, 20922100.Google Scholar
Reddon, AR and Hurd, PL (2013) Water pH during early development influences sex ratio and male morph in a West African cichlid fish, Pelvicachromis pulcher. Zoology 116, 139143.Google Scholar
Rawson, PD and Burton, RS (2002) Functional coadaptation between cytochrome c and cytochrome c oxidase within allopatric populations of a marine copepod. Proceedings of the National Academy of Sciences of the United States of America 99, 1295512958.Google Scholar
Reis, RE (2003) Family Callichthyidae. In Reis, RE, Kullander, SO and Ferraris, CJ Jr (eds). Checklist of the Freshwater Fishes of South and Central America. Porto Alegre, Brazil: Edipucrs, pp. 291309.Google Scholar
Ribeiro, AC (2006) Tectonic history and the biogeography of the freshwater fishes from the coastal drainages of eastern Brazil: an example of faunal evolution associated with a divergent continental margin. Neotropical Ichthyology 4, 225246.Google Scholar
Robertsen, G, Hansen, H, Bachmann, L and Bakke, TA (2007) Arctic charr (Salvelinus alpinus) is a suitable host for Gyrodactylus salaris (Monogenea, Gyrodactylidae) in Norway. Parasitology 134, 257267.Google Scholar
Robinson, BW and Wilson, DS (1994) Character release and displacement in fishes: a neglected literature. The American Naturalist 144, 596627.Google Scholar
Rodrigues, R, Schneider, H, Santos, S, Vallinoto, M, Sain-Paul, U and Sampaio, I (2008). Low levels of genetic diversity depicted from mitochondrial DNA sequences in a heavily exploited marine fish (Cynoscion acoupa, Sciaenidae) from the Northern coast of Brazil. Genetics and Molecular Biology 31, 487492.Google Scholar
Rogers, AR and Harpending, H (1992) Population growth makes waves in the distribution of pairwise genetic differences. Molecular Biology and Evolution 9, 552569.Google Scholar
Satoh, TP, Miya, M, Mabuchi, K and Nishida, M (2016) Structure and variation of the mitochondrial genome of fishes. BMC Genomics 17, 120. http://dx.doi.org/10.1186/s12864-016-3054-y.Google Scholar
Sbisà, E, Tanzariello, F, Reyes, A, Pesole, G and Saccone, C (1997) Mammalian mitochondrial D-loop region structural analysis: identification of new conserved sequences and their functional and evolutionary implications. Gene 205, 125140.Google Scholar
Seehausen, O (2006) African cichlid fish: a model system in adaptive radiation research. Proceedings of the Royal Society of London B 273, 19871998.Google Scholar
Shen, Y-Y, Liang, L, Zhu, Z-H, Zhou, W-P, Irwin, DM and Zhang, Y-P (2010) Adaptive evolution of energy metabolism genes and the origin of flight in bats. Proceedings of the National Academy of Sciences of the United States of America 107, 86668671.Google Scholar
Shibatta, OA and Hoffmann, AC (2005) Variação geográfica em Corydoras paleatus (Jenyns) (Siluriformes, Callichthyidae) do sul do Brasil. Revista Brasileira de Zoologia 22, 366371.Google Scholar
Shinn, AP, Gibson, DI and Sommerville, C (2001) Morphometric discrimination of Gyrodactylus salaris Malmberg (Monogenea) from species of Gyrodactylus parasitising British salmonids using novel parameters. Journal of Fish Diseases 24, 8397.Google Scholar
Siddall, M, Rohling, EJ, Almogi-Labin, A, Hemleben, Ch, Meischner, D, Schmelzer, I and Smeed, DA (2003) Sea-level fluctuations during the last glacial cycle. Nature 423, 853858.Google Scholar
Smith, TB and Skúlason, S (1996) Evolutionary significance of resource polymorphisms in fishes, amphibians, and birds. Annual Review of Ecology and Systematics 27, 111133.Google Scholar
Solomon, SG, Okomoda, VT and Ogbenyikwu, AI (2015) Intraspecific morphological variation between cultured and wild Clarias gariepinus (Burchell) (Clariidae, Siluriformes). Archives of Polish Fisheries 23, 5361.Google Scholar
Staden, R (1996) The Staden sequence analysis package. Molecular Biotechnology 5, 233241.Google Scholar
Stern, A, Doron-Faigenboim, A, Erez, E, Martz, E, Bacharach, E and Pupko, T (2007) Selecton 2007: advanced models for detecting positive and purifying selection using a Bayesian inference approach. Nucleic Acids Research 35, W506W511.Google Scholar
Storz, JF (2005) Using genome scans of DNA polymorphism to infer adaptive population divergence. Molecular Ecology 14, 671688.Google Scholar
Strauss, RE (1985) Evolutionary allometry and variation in body form in the South American catfish genus Corydoras (callichthyidae). Systematic Zoology 34, 381396.Google Scholar
Suguio, K and Martin, L (1978) Quaternary marine formations of the state of São Paulo and southern Rio de Janeiro. 1978 International Symposium on Coastal Evolution in the Quaternary, Special Publication, No. 1, pp. 155.Google Scholar
Swanson, WJ, Nielsen, R and Yang, Q (2003) Pervasive adaptive evolution in mammalian fertilization proteins. Molecular Biology and Evolution 20, 1820.Google Scholar
Tajima, F (1989) Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics 123, 585595.Google Scholar
Takahashi, T and Koblmüller, S (2011) The adaptive radiation of cichlid fish in Lake Tanganyika: a morphological perspective. International Journal of Evolutionary Biology 2011, 115. http://dx.doi.org/10.4061/2011/620754.Google Scholar
Tamura, K, Stecher, G, Peterson, D, Filipski, A and Kumar, S (2013) MEGA6: molecular evolutionary genetics analysis version 6.0. Molecular Biology and Evolution 30, 27252729.Google Scholar
Thompson, JN (1999) Specific hypotheses on the geographic mosaic of coevolution. The American Naturalist 153, S1S14.Google Scholar
Thompson, JN (2005) The Geographic Mosaic of Coevolution. Chicago, USA: The University of Chicago Press.Google Scholar
Trippel, EA and Harvey, HH (1987) Abundance, growth, and food supply of white suckers (Catostomus commersoni) in relation to lake morphometry and pH. Canadian Journal of Zoology 65, 558564.Google Scholar
Tschá, MK (2016) Taxon Pulse: Um modelo para a diversificação genética de peixes em bacias hidrográficas costeiras (Doctoral thesis). UFPR, Curitiba, Brazil.Google Scholar
Tschá, MK, Baggio, RA, Marteleto, FM, Abilhoa, V, Bachmann, L and Boeger, WA (2017) Sea-level variations have influenced the demographic history of estuarine and freshwater fishes of the coastal plain of Paraná, Brazil. Journal of Fish Biology 90, 968979.Google Scholar
Turan, C, Yalçin, S, Turan, F, Okur, E and Akyurt, İ (2005) Morphometric comparisons of African catfish, Clarias gariepinus, populations in Turkey. Folia Zoologica 54, 165172.Google Scholar
von der Heyden, S, Lipinski, MR and Matthee, CA (2010) Remarkably low mtDNA control region diversity in an abundant demersal fish. Molecular Phylogenetics and Evolution 55, 11831188.Google Scholar
Walter, RP, Blum, MJ, Snider, SB, Paterson, IG, Bentzen, P, Lamphere, BA and Gilliam, JF (2011) Isolation and differentiation of Rivulus hartii across Trinidad and neighboring islands. Molecular Ecology 20, 601618.Google Scholar
Watterson, GA (1978) The homozygosity test of neutrality. Genetics 88, 405417.Google Scholar
Wolff, LL and Hahn, NS (2017) Fish habitat associations along a longitudinal gradient in a preserved coastal Atlantic stream, Brazil. Zoologia 34, 113. http://dx.doi.org/10.3897/zoologia.34.e12975.Google Scholar
Wright, S (1943) Isolation by distance. Genetics 28, 114138.Google Scholar
Xavier, R, Faria, PJ, Paladini, G, van Oosterhout, C, Johnson, M and Cable, J (2015) Evidence for cryptic speciation in directly transmitted gyrodactylid parasites of Trinidadian guppies. PLoS ONE 10, 115. http://dx.doi.org/10.1371/journal.pone.0117096.Google Scholar
Yang, Z, Nielsen, R, Goldman, N and Pedersen, A-MK (2000) Codon-substitution models for heterogeneous selection pressure at amino acid sites. Genetics 155, 431449.Google Scholar
Zeng, K, Shi, S and Wu, C-I (2007) Compound tests for the detection of hitchhiking under positive selection. Molecular Biology and Evolution 24, 18981908.Google Scholar
Zhang, J, Cai, Z and Huang, L (2006) Population genetic structure of crimson snapper Lutjanus erythropterus in East Asia, revealed by analysis of the mitochondrial control region. ICES Journal of Marine Science 63, 693704.Google Scholar
Ziętara, MS and Lumme, J (2002) Speciation by host switch and adaptive radiation in a fish parasite genus Gyrodactylus (monogenea, Gyrodactylidae). Evolution 56, 24452458.Google Scholar
Ziętara, MS, Arndt, A, Geets, A, Hellemans, B and Volckaert, FAM (2000) The nuclear rDNA region of Gyrodactylus arcuatus and G. branchicus (monogenea: Gyrodactylidae). Journal of Parasitology 86, 13681373.Google Scholar
Supplementary material: PDF

Bueno-Silva and Boeger supplementary material

Figure S1

Download Bueno-Silva and Boeger supplementary material(PDF)
PDF 298.2 KB