Research Article
Print
Research Article
Molecular perspective on the American transisthmian species of Macrobrachium (Caridea, Palaemonidae)
expand article infoLeonardo Pileggi, Natalia Rossi§, Ingo S. Wehrtmann|, Fernando L Mantelatto
‡ 2Laboratory of Bioecology and Crustacean Systematics, Faculty of Philosophy, Sciences and Letters at Ribeirão Preto (FFCLRP), University of São Paulo, Ribeirão Preto, São Paulo, Brazil
§ University of São Paulo - FFCLRP, Ribeirão Preto, São Paulo, Brazil
| Universidad de Costa Rica, San Pedro de Montes de Oca, Costa Rica
¶ University of Sao Paulo (USP), Ribeirão Preto - São Paulo, Brazil
Open Access

Abstract

The closure of the Isthmus of Panama (about 3.1 million years ago) separated previously continuous populations and created two groups of extant species, which live now in the Pacific and Atlantic drainage systems. This relatively recent event was a trigger to diversification of various species in the Neotropics, nonetheless there are exemplars that do not show sufficient morphologic variability to separate them by traditional morphological tools. About 60 years ago, some freshwater decapod species with high morphological similarity were separate by previous researchers, based on geographical distribution, in Pacific and Atlantic and considered as “sister species”. However, the complete isolation of these prawns by this geographical barrier is questionable, and it has generated doubts about the status of the following transisthmian pairs of sibling species: Macrobrachium occidentale × M. heterochirus, M. americanum × M. carcinus, M. digueti × M. olfersii, M. hancocki × M. crenulatum, M. tenellum × M. acanthurus and M. panamense × M. amazonicum. Here we evaluated the relation among these pairs of sibling species in a molecular phylogenetic context. We generated 95 new sequences: 26 sequences of 16S rDNA, 25 of COI mtDNA and 44 of 18S nDNA. In total, 181 sequences were analyzed by maximum likelihood phylogenetic method, including 12 Macrobrachium transisthmian species, as well as seven other American Macrobrachium species, and two other palaemonids. Our analysis corroborated the morphological proximity of the sibling species. Despite the high degree of morphological similarities and considerable genetic diversification encountered among the transisthmian sister species, our data support the conclusion that all species included in sibling groups studied herein are valid taxonomic entities, but not all pairs of siblings form natural groups.

Keywords

Freshwater decapods, genetic variability, molecular phylogeny, Palaemoninae , sibling species

Introduction

In the late Pliocene, the closure of the Isthmus of Panama was a trigger to the diversification of many species in the Neotropics. The separation of previously continuous populations created two groups of extant species, which live now in the Atlantic and Pacific drainage systems. This vicariant event opened a unique opportunity for studies on evolution, divergence and speciation processes (Knowlton et al. 1993, Knowlton and Weigt 1998, Lessios 2008). The Central American land bridge is a well-dated biogeographic barrier and is a relatively recent event, about 3.1 million years ago (Keigwin 1978, Coates et al. 1992, Coates and Obando 1996, Anger 2013). Since then, the Atlantic and Pacific marine ecosystems became gradually separated, whereas the gene flow was blocked between organisms on either side.

In spite of the geographic separation, some species are difficult or impossible to distinguish using traditional morphological features, and are thus called “sibling species” (see Knowlton 1993 and references cited therein). These sibling species refer to pairs of species that are genetically closely related, but reproductively isolated (Mayr 1963, Steyskal 1972, Knowlton 1986). Others authors refer to “sibling” as “geminate species” (Jordan 1908, Stillman and Reeb 2001, Marko 2002), in which individuals were separated necessarily by a geographic barrier, and each member of the pair occurs along one coast of the Americas (Lessios 1998, Miura et al. 2010). Other non-morphological features have been used to distinguish these species such as “karyology, hybridization experiments to detect postzygotic incompatibility, distribution patterns, resource use, breeding season, life history and development, mating behavior (including visual, acoustical, and chemical signals), color pattern, and various biochemical characters” (Knowlton 1986). Consequently, a pair of species, reproductively isolated and very similar in morphology, is not necessarily considered as sibling species, and an interdisciplinary approach is necessary to evaluate this conclusion.

Molecular tools have been used to contribute with species delimitation in several cryptic decapods (Schubart et al. 2001a, b, Kitaura et al. 2002, Lai et al. 2010, Pileggi and Mantelatto 2010, Mantelatto et al. 2011, Negri et al. 2012, Torati and Mantelatto 2012). Phylogenies based on molecular data has evidenced probable cases of misidentification of sibling species based on morphology (Lessios 2008, Rossi and Mantelatto 2013). For some freshwater species, the isolation by the closure of the Isthmus of Panama might be questionable, since species of the genus Macrobrachium Spence Bate, 1868 can disperse over greater distances than the width of the Isthmus (Steeves et al. 2005, Bauer and Delahoussaye 2008, Bauer 2011) and may also use the Panama Canal as passageway for both sides (Hildebrand 1939, Abele and Kim 1989).

Most studies on decapods sister species focused only in marine species of the genus Alpheus Fabricius, 1798 (Knowlton et al. 1993, Knowlton and Weigt 1998, Wehrtmann and Albornoz 2002), while our knowledge of the impact of the Isthmus of Panama on freshwater-invading decapods is extremely limited (Anger 2013). Prawns of the genus Macrobrachium are widely distributed in rivers of tropical and subtropical regions with more than 240 recognized species worldwide (De Grave and Fransen 2011). Although its greatest diversity has been found in the Indo-Pacific region, in the Americas there are more than 55 valid species, representing an area of great importance concerning the diversity of the family Palaemonidae (Holthuis 1952, Pileggi and Mantelatto 2012).

The high morphological similarity between some American species led Holthuis (1952) to designate Atlantic and Pacific Macrobrachium “sister species”. Until now, morphological similarities between the transisthmian “sibling species” have impeded the identification of the following pairs of species: Macrobrachium occidentale × M. heterochirus, M. americanum × M. carcinus, M. digueti × M. olfersii, M. hancocki × M. crenulatum, M. tenellum × M. acanthurus and M. panamense × M. amazonicum. These species occur primarily in Central America, with the first species of each pair is found in the Pacific drainage and the second in the Atlantic side. Larvae of these species require saline water (i.e., 10–35 ppt) to complete their life cycle, and exhibit other adaptive features, such as extended larval development and amphidromous life histories (Hedgpeth 1949, Bauer and Delahoussaye 2008, Bauer 2011, 2013). Moreover, these prawns show great morphological modifications during ontogenesis, and as other congeneric species they present controversial systematic issues, with high interspecific conservatism and males with intraspecific variation, as found among distinct morphotypes (Holthuis 1952, Moraes-Riodades and Valenti 2004, Pileggi and Mantelatto 2010, Vergamini et al. 2011). Considering the doubt whether the previously indicated species of Central American Macrobrachium are sister taxa or not, our study aimed to evaluate in a molecular phylogenetic context the relationships among 12 transisthmian Macrobrachium “sibling species” from the Americas in order to assess the validity of their current species level.

Methods

Sample collection

Fresh specimens for molecular analysis were obtained from field collections in rivers and estuaries in Brazil, Chile, Venezuela, and Costa Rica (Table 1). The individuals were preserved in 75–90% ethanol and deposited in the Crustacean Collection of the Department of Biology (CCDB), Faculty of Philosophy, Sciences and Letters at Ribeirão Preto (FFCLRP), University of São Paulo (USP), National Institute of Research of Amazônia (INPA) – Brazil, and the Museum of Zoology, School of Biology, University of Costa Rica, Costa Rica (MZUCR). The collections of species conducted in this study complied with current applicable state and federal laws.

Table 1.

Trans-isthmian species of Macrobrachium and other palaemonids used for the phylogenetic analyses, with the respective collection locality, distribution of the species, catalogue number, and genetic database accession numbers at GenBank.

Species Locality Distribution Catalogue Nº 16S COI 18S
Sibling species of Macrobrachium
M. acanthurus -1 Ilha de São Sebastião-SP, Brazil America-Atlantic CCDB 2134 HM352445 HM352485 KM101492
M. acanthurus -2 Guaraqueçaba-PR, Brazil America-Atlantic CCDB 2546 HM352444 KM101538 KM101493
M. acanthurus -1 Puerto Viejo, Costa Rica America-Atlantic CCDB 1556 KM101464 KM101537 KM101491
M. acanthurus -2 Cahuita, Costa Rica America-Atlantic CCDB 2901 KM101465 KM101539 KM101494
M. acanthurus -1 Bocas del Toro, Panama America-Atlantic CCDB 3538 KM101467 KM101541 KM101496
M. acanthurus -2 Panama America-Atlantic CCDB 3536 KM101466 KM101540 KM101495
M. tenellum -1 Puntarenas, Costa Rica North/Central America-Pacific MZUCR 1936-002 KM101488 KM101567 KM101534
M. tenellum -2 Guanacaste, Costa Rica North/Central America-Pacific MZUCR 3290-01 KM101489 KM101568 KM101535
M. tenellum Oaxaca, Mexico North/Central America-Pacific CNCR 24831 KM101487 KM101566 KM101533
M. amazonicum -1 Santana-AP, Brazil South/Central America-Atlantic CCDB 1965 HM352441 HM352486 KM101497
M. amazonicum -2 Aquidauana-MS, Brazil South/Central America-Atlantic CCDB 1970 HM352442 HM352487 -
M. amazonicum -3 Itacoatiara-AM, Brazil South/Central America-Atlantic CCDB 2085 HM352443 HM352488 -
M. amazonicum Panama South/Central America-Atlantic CNCR 5151 KM101468 KM101542 KM101498
M. panamense -1 Cerca Camaronera, Costa Rica Central America-Pacific MZUCR 2972-01 KM101485 KM101562 KM101528
M. panamense -2 Río Tempisque, Costa Rica Central America-Pacific MZUCR 2971-01 KM101484 KM101561 KM101527
M. panamense -3 Guanacaste, Costa Rica Central America-Pacific MZUCR 3291-01 KM101486 KM101563 KM101529
M. olfersii -1 Ilha de São Sebastião-SP, Brazil America-Atlantic CCDB 2435 HM352459 HM352496 KM101523
M. olfersii -2 Antonina-PR, Brazil America-Atlantic CCDB 2445 HM352458 KM101558 KM101524
M. olfersii Isla Margarita, Venezuela America-Atlantic CCDB 2446 HM352460 KM101559 KM101525
M. olfersii -1 Reserva Veragua, Costa Rica America-Atlantic CCDB 4873 KM101483 KM101560 KM101526
M. olfersii -2 Costa Rica (Atlantic) America-Atlantic CCDB 2876 JQ805835 JQ805933 JQ805858
M. olfersii -3 Costa Rica (Atlantic) America-Atlantic CCDB 2880 JQ805839 JQ805936 JQ805859
M. digueti -1 Costa Rica (Pacific) South/Central America-Pacific CCDB 2882 JQ805806 JQ805903 JQ805847
M. digueti -2 Costa Rica (Pacific) South/Central America-Pacific CCDB 3091 JQ805807 JQ805904 -
M. digueti -3 Río Aranjuez, Costa Rica Central America-Pacific MZUCR 3292-01 KM101476 KM101551 KM101514
M. digueti Mexico South/Central America-Pacific CNCR 24811 JQ805808 JQ805906 JQ805849
M. crenulatum -1 Isla Margarita, Venezuela South/Central America-Atlantic CCDB 2124 HM352463 HM352498 KM101512
M. crenulatum -2 Venezuela South/Central America-Atlantic IVIC 123 JQ805801 - JQ805845
M. crenulatum -1 Costa Rica South/Central America-Atlantic CCDB 2873 JQ805804 JQ805900 JQ805846
M. crenulatum -2 Costa Rica South/Central America-Atlantic CCDB 2877 JQ805800 - JQ805844
M. crenulatum -3 Reserva Veragua, Costa Rica South/Central America-Atlantic CCDB 4874 KM101475 KM101550 KM101513
M. hancocki -1 Costa Rica South/Central America-Pacific CCDB 3090 JQ805813 JQ805911 -
M. hancocki -2 Costa Rica South/Central America-Pacific CCDB 3092 JQ805814 JQ805912 JQ805851
M. hancocki -3 Costa Rica South/Central America-Pacific CCDB 3757 JQ805821 JQ805920 -
M. hancocki -4 Costa Rica South/Central America-Pacific CCDB 3756 JQ805822 JQ805919 -
M. hancocki Panama South/Central America-Pacific RMNHD 8810 JQ805817 JQ805915 JQ805852
M. carcinus -1 Santana-AP, Brazil America-Atlantic CCDB 2122 HM352448 HM352490 KM101507
M. carcinus -2 Ubatuba-SP, Brazil America-Atlantic CCDB 2136 HM352449 HM352491 KM101509
M. carcinus Isla Margarita, Venezuela America-Atlantic CCDB 2123 HM352450 HM352492 KM101508
M. carcinus -1 Río Suarez, Costa Rica America-Atlantic CCDB 2145 HM352452 KM101548 KM101510
M. carcinus -2 Cahuita, Costa Rica America-Atlantic CCDB 4876 KM101474 KM101549 KM101511
M. americanum -1 Costa Rica South/Central America-Pacific CCDB 1731 HM352447 HM352489 KM101499
M. americanum -2 Río Aranjuez, Costa Rica South/Central America-Pacific MZUCR 3292-03 KM101473 KM101547 KM101504
M. americanum -3 Río Coronado, Costa Rica South/Central America-Pacific MZUCR 2963-01 KM101470 KM101544 KM101501
M. americanum -4 Río Oro, Costa Rica South/Central America-Pacific MZUCR 2964-01 KM101471 KM101545 KM101502
M. americanum -5 Isla Violines, Costa Rica South/Central America-Pacific MZUCR 2970-01 KM101472 KM101546 KM101503
M. americanum -6 Costa Rica South/Central America-Pacific CCDB 2883 JQ805797 JQ805899 JQ805843
M. americanum Río Cabuya, Panama South/Central America-Pacific CCDB 2463 KM101469 KM101543 KM101500
M. heterochirus Ilha de São Sebastião-SP, Brazil South/Central America-Atlantic CCDB 2137 HM352454 HM352494 KM101515
M. heterochirus -1 Río Suarez, Costa Rica South/Central America-Atlantic CCDB 2899 KM101477 KM101552 KM101516
M. heterochirus -2 Reserva Veragua, Costa Rica South/Central America-Atlantic CCDB 4875 KM101478 KM101553 KM101517
M. heterochirus Veracruz, Mexico South/Central America-Atlantic Not available KM101479 KM101554 KM101518
M. occidentale Río Aranjuez, Costa Rica North/Central America-Pacific MZUCR 3292-02 KM101482 KM101557 KM101522
M. occidentale Oaxaca, Mexico North/Central America-Pacific CNCR 24838 KM101481 KM101556 KM101521
Other palaemonids
M. borellii Buenos Aires, Argentina South America-Inland waters UFRGS 3669 HM352426 HM352480 KM101505
M. brasiliense Serra Azul-SP, Brazil South America-Inland waters CCDB 2135 HM352429 HM352481 KM101506
M. jelskii Pereira Barreto-SP, Brazil South America-Inland waters CCDB 2129 HM352437 HM352484 KM101519
M. michoacanus Oaxaca, Mexico Mexico-Inland waters CNCR 24837 KM101480 KM101555 KM101520
M. potiuna Eldorado-SP, Brazil Brazil-Inland waters CCDB 2131 HM352438 KM101564 KM101530
M. rosenbergii Culture, Brazil Indo-Pacific CCDB 2139 HM352465 - KM101531
M. rosenbergii Kaohsiung Co., Taiwan Indo-Pacific Not informed - AB235295 -
M. surinamicum Icangui-PA, Brazil South America-Atlantic INPA-CR 183 HM352446 KM101565 KM101532
Cryphiops caementarius Region IV, Chile South America-Pacific CCDB 1870 HM352453 HM352495 KM101490
Palaemonetes argentinus Parati-RJ, Brazil South America CCDB 2011 HM352425 - KM101536
Palaemonetes argentinus Not informed South America Not informed - HQ587179 -

Additional material was obtained by donation, visit or loan from distinct worldwide crustacean collections (Table 1). A total of 65 specimens of Macrobrachium and three of other genera were analyzed. Almost all sequences were generated in the Laboratory of Bioecology and Crustacean Systematics (LBSC). Some additional comparative sequences were retrieved from GenBank (Table 1). The selection of the other Macrobrachium species and genera was based on the phylogeny of Pileggi and Mantelatto (2010), including closely related as well as more phylogenetically distant species. The species identification was based on diagnostic morphological features in accordance with the literature (Holthuis 1952, Villalobos 1969, Melo 2003, Pileggi and Mantelatto 2012).

DNA extraction, amplification and sequencing

The molecular analysis was based on partial fragments of the 16S rDNA, 18S nDNA and COI mtDNA genes, which have been effective in solving different levels of relationships among decapod species (Schubart et al. 2000, 2001a, b, Porter et al. 2005, Pileggi and Mantelatto 2010, Mantelatto et al. 2011, Vergamini et al. 2011, Carvalho et al. 2013, Rossi and Mantelatto 2013).

DNA extraction, amplification and sequencing protocols followed Pileggi and Mantelatto (2010). Total genomic DNA was extracted from the muscle tissue of walking legs, the chelipeds, or the abdomen. An approximately 530-bp region of the 16S rDNA gene, 560-bp region of the COI gene and 550-bp region of the nuclear 18S gene were amplified from diluted DNA by means of a polymerase chain reaction (PCR) in an Applied Biosystems Veriti 96 Well Thermal Cycler® (thermal cycles: initial denaturing for 5 min at 95 °C; annealing for 40 cycles: 45s at 95 °C, 45s at 48–50 °C, 1 min at 72 °C; final extension 3 min at 72 °C) with the following primers: 16Sar and 16Sbr (Palumbi et al. 1991) for 16S mitochondrial gene; COI-a and COI-f (Palumbi and Benzie 1991) for COI mitochondrial gene; 18Sai and 18Sb3.0 (Whiting et al. 1997) for 18S nuclear gene. PCR products were purified using Sure Clean (Bioline) and sequenced with the ABI Big Dye® Terminator Mix (Applied Biosystems, Carlsbad, CA) in an ABI Prism 3100 Genetic Analyzer® (Applied Biosystems automated sequencer) following Applied Biosystems protocols. All sequences were confirmed by sequencing both strands. A consensus sequence for the two strands was obtained using the computational program BIOEDIT 7.0.5 (Hall 2005). Apart from that, the consensus sequences were blasted on GenBank and compared with our previous sequences. Genetic vouchers generated were deposited in the CCDB under the accession numbers listed in Table 1 or returned with an appropriate label to the original collections.

Molecular analyses

Sequences were aligned using CLUSTAL W (Thompson et al. 1994) with interface to BIOEDIT with default parameters. Ambiguous alignment regions were removed. Genetic-distance analyzes for the partial sequences of the three markers (16S rDNA, COI mtDNA and 18S nDNA), over sequence pairs between and within groups were conducted in MEGA 5.2 using Kimura-2-parameter model (Tamura et al. 2011). Sequences were analyzed under the Akaike Information Criterion (AIC) (Posada and Buckley 2004) with the program JMODELTEST 2.1.3 (Darriba et al. 2012) to find the best substitution model. The maximum likelihood (ML) analysis was carried out using PAUP 4.0b10 (Swofford 2003). The consistency of topologies was measured by the bootstrap method (1000 replicates), and only confidence values > 50% were reported.

Results

Our phylogenetic analysis included 12 transisthmian American species of Macrobrachium, 7 from other American Macrobrachium species, and 2 from palaemonid-related groups. We generated 95 new sequences: 26 mitochondrial 16S sequences, 25 mitochondrial COI sequences, and 44 nuclear 18S sequences. The analysis of the 181 sequences from the three genes produced an alignment of 1.645 bp.

The optimal model for the concatenated data set was the TPM1uf model of sequence evolution (Kimura 1981) plus gamma distributed rate heterogeneity with a significant proportion of invariable sites (TPM1uf +I+G) with the following parameters: assumed nucleotide frequencies A = 0.3028, C = 0.2125, G = 0.1909, T = 0.2937; proportion of invariable sites I = 0.6020; the variable sites followed a gamma distribution, with shape parameter = 0.6700.

The topology obtained by maximum likelihood from concatenated genes (16S, 18S and COI) analyses confirmed that the transisthmian sibling species (M. heterochirus × M. occidentale – Sibling 1, M. carcinus × M. americanum – Sibling 2, M. olfersii × M. digueti – Sibling 3, M. crenulatum × M. hancocki – Sibling 4, and M. acanthurus × M. tenellum – Sibling 5) are closely related by well-supported clades (Fig. 1). Sibling 6 (M. amazonicum × M. panamense) did not form a separate sister clade despite being phylogenetically close. The position of Palaemonetes argentinus showed a stable condition in an external branch. However, the other outgroup (Cryphiops caementarius) was maintained within the Macrobrachium clade in the phylogeny (Fig. 1). The results did not reveal geographical separation among populations of the same species inside each group (Siblings 1–5). Macrobrachium michoacanus (see the arrow in the phylogeny) seems to be close related to M. hancocki in Sibling 4 group.

Figure 1. 

Phylogenetic tree obtained from concatenated maximum likelihood analysis of 16S, COI and 18S sequences for Macrobrachium sibling species. Numbers are significance values for 1000 bootstraps; values ≤ 50% are not shown. Abbreviations: ARG: Argentina; BR: Brazil; CH: Chile; CR: Costa Rica; MX: Mexico; PN: Panama; VZ: Venezuela. A: lateral view of the rostrum of M. amazonicum; B: lateral view of the rostrum of M. olfersii. C: lateral view of the rostrum of M. carcinus.

The relation among the sibling groups is supported by morphological traits. The species included in Siblings 1 and 2 exhibit similar shapes of the rostrum with the upper margin somewhat arched over the eye and with the apex directed upward (Fig. 1C). Species of the Siblings 3 and 4 with M. michoacanus and M. surinamicum show similar rostrum, being almost straight and usually with more than 10 teeth in the upper margin (Fig. 1B). In the same way the species of Siblings 5 and 6 and M. jelskii have a distinct rostrum, which is elongated, slender, with apex curved upward, with many teeth in the upper and lower margin (Fig. 1A).

In general, distance analyses revealed that the percentage of intraspecific variation was lower than interspecific variation (Table 2). Considering the relation between distinct sibling species, the genetic variability ranged from 4.4% (Sibling 3 × Sibling 4) to 16.9% (Sibling 4 × Sibling 6) for 16S, from 11.3% (Sibling 3 × Sibling 4) to 23.9% (Sibling 2 × Sibling 5) for COI, and from 1.1% (Sibling 5 × Sibling 6) to 11.3% (Sibling 2 × Sibling 5, 6) for 18S (Table 2). Inside each sibling group, the genetic variability varied between 1.5% (Sibling 3) and 8.7% (Sibling 6) for 16S, between 8.3% (Sibling 2) and 16.9% (Sibling 5) for COI, and between 0.0% (Sibling 2, 3) and 1.1% (Sibling 1) for 18S (Table 2).

Table 2.

Genetic divergence matrix of the 16S and COI mitochondrial genes and 18S nuclear gene among American Macrobrachium sibling species obtained by distance analyses using Kimura-2-parameter model. SB: Sibling species. Comparison between the same sibling (bold numbers) comprises interspecific and intraspecific (numbers in parenthesis) analyses.

SB1 SB2 SB3 SB4 SB5 SB6
16S SB1 0.047–0.046 (0.002–0.013)
SB2 0.088–0.103 0.019–0.028 (0.000–0.006)
SB3 0.076–0.093 0.084–0.102 0.015–0.019 (0.000–0.004)
SB4 0.081–0.097 0.076–0.098 0.044–0.065 0.017–0.021 (0.000–0.011)
SB5 0.095–0.136 0.107–0.125 0.115–0.128 0.117–0.136 0.064–0.069 (0.000–0.004)
SB6 0.112–0.146 0.114–0.149 0.115–0.155 0.117–0.169 0.062–0.097 0.075–0.087 (0.002–0.011)
COI SB1 0.110–0.128 (0.011–0.061)
SB2 0.175–0.233 0.083–0.122 (0.000–0.038)
SB3 0.149–0.179 0.159–0.204 0.103–0.119 (0.004–0.022)
SB4 0.136–0.179 0.168–0.205 0.113–0.168 0.086–0.109 (0.006–0.091)
SB5 0.156–0.197 0.167–0.239 0.147–0.191 0.168–0.209 0.160–0.169 (0.000–0.022)
SB6 0.151–0.180 0.161–0.234 0.143–0.190 0.148–0.196 0.138–0.187 0.141–0.152 (0.004–0.040)
18S SB1 0.011 (0.000)
SB2 0.097–0.100 0.000 (0.000)
SB3 0.059–0.097 0.097 0.000 (0.000)
SB4 0.044–0.097 0.094–0.097 0.022–0.025 0.008 (0.000)
SB5 0.056–0.059 0.110–0.113 0.053–0.056 0.041–0.047 0.003 (0.000)
SB6 0.056–0.061 0.103–0.113 0.047–0.056 0.039–0.047 0.000–0.011 0.008 (0.000)

Discussion

Over 150 sequences from three different gene regions were used in the present study in order to estimate phylogenetic relationships among freshwater prawns of the genus Macrobrachium, which previously were assumed to be transisthmian sibling species. The results revealed that all geminate species studied herein were valid taxonomic entities. Likewise they confirmed the role of the Isthmus of Panama as an effective barrier contributing in the separation of sibling species by the mechanism of allopatric speciation. However, in other cases the separation happened before the closure of the Isthmus probably by the mechanism of sympatric speciation. Our multigenic phylogeny produced consistent groups in most of the pairs of geminate species i.e., sister taxa geographically separated: Macrobrachium heterochirus × M. occidentale, M. carcinus × M. americanum, M. olfersii × M. digueti, M. crenulatum × M. hancocki and M. acanthurus × M. tenellum. The constitution of these clades corroborates the morphological proximity of each pair of species as mentioned by Holthuis (1952).

The genetic divergence analyses showed the separation of each sibling group from others, which suggest a consistent relation in comparison with other congeners species (Table 2). Considering that for 16S the divergence in decapods is presumed to be around between 0.6 to 0.9% per Myr (Schubart et al. 2000), we can estimate the divergence time of the sibling species according to the closure of the Isthmus. For Siblings 1 and 5 the time of divergence between the species was approximately from 5.1 to 7.8 and 7.11 to 11.5 Mya, respectively. These estimates predate the closure of the Isthmus, which suggest that the speciation process separated these two species already before the closure. Considering that these amphidromous species are dependent of estuarine water for successful larval development, a sympatric speciation hypothesis seems to be unlikely. However, in these cases the possibility of occurrence of this event is plausible, probably due to environmental changes (Knowlton et al. 1993, Knowlton and Weigt 1998, Morrison et al. 2004). A genetic differentiation could have arisen from a mutational step and the two subpopulations, whose geographic distribution ranges overlap completely, became isolated because both occupy completely different ecological niches (Bush 1994). Analogous events were reported for other crustacean (Malay and Paulay 2009, Jarman et al. 2011). In addition, the estuary can contribute to restriction of the gene flow between the species by distinct selective regimes or habitat fidelity of the species, generating potential speciation in complete or partial isolation (Stanhope et al. 1992, Bilton et al. 2002). Therefore, the sympatric speciation may have occurred in these sibling species by the mechanism of microallopatry (Fitzpatrick et al. 2008). The difficulty in separating M. heterochirus from M. occidentale, and M. acanthurus from M. tenellum using morphological, ecological and genetic characters (Tables 2, 3, 5), allied with the consistent position in the phylogeny (Fig. 1) provide convincing arguments to consider them as sibling species. The phylogenetic position of Siblings 1 with 2 and Siblings 5 with “6” (here marked between quotes due its artificial position, not characterized as sibling) followed the morphological grouping based on the shape of the rostrum (Fig. 1A, C) indicating that this character is determinant for taxonomic studies.

Table 3.

Distributional and ecological comparison among each Macrobrachium species of sibling pair 1 and 2.

Sibling 1 Sibling 2
M. occidentale M. heterochirus M. americanum M. carcinus
American slope Pacific Atlantic Pacific Atlantic
Distribution Mexico to Panama USA (Florida) to Brazil (Rio Grande do Sul) Mexico (Baja California) to Peru USA (Florida) to Brazil (Rio Grande do Sul)
Habitat wide range of altitudes (more common in higher elevations of the rivers) wide range of altitudes (more common in medium and higher courses of the rivers
Reproduction require brackish water for reproduction (extended larval development with numerous and small eggs) require brackish water for reproduction (extended larval development with numerous and small eggs)
Morphology very similar and just a few morphological details better seen in adult males are useful characters to separate both species very similar and present few distinct morphological characters
References Holthuis 1952, Mejía-Ortiz et al. 2001, Rocha and Bueno 2004, Almeida et al. 2008, Lara 2009, Villalobos-Hiriart et al. 2010, Lara and Wehrtmann 2011, García-Guerrero et al. 2013, Pileggi et al. 2013 Holthuis 1952, Choudhury 1971, Monaco 1975, Bowles et al. 2000, Mejía-Ortíz et al. 2001, Hernández et al. 2007, Valencia and Campos 2007, Almeida et al. 2008, Lara 2009, Pileggi and Mantelatto 2010, Lara and Wehrtmann 2011, García-Guerrero et al. 2013
Table 4.

Distributional and ecological comparison among each Macrobrachium species of sibling pair 3 and 4.

Sibling 3 Sibling 4
M. digueti M. olfersii M. hancocki M. crenulatum
American slope Pacific Atlantic Pacific Atlantic
Distribution Mexico (Baja California) to Ecuador USA (Florida) to Brazil (Rio Grande do Sul) Costa Rica to Ecuador West Indies, Panama, Colombia and Venezuela
Habitat wide range of altitudes (more common in higher elevations of the rivers) wide range of altitudes (more common in higher elevations of the rivers)
Ecology require brackish water for reproduction (extended larval development with numerous and small eggs) require brackish water for reproduction (extended larval development with numerous and small eggs)
Morphology very alike a few characters better seen in adult males are used to separate both species very similar and can be differentiated using the color pattern, but fixed specimens are difficult to distinguish using only morphological characters
References Holthuis 1952, Dugger and Dobkin 1975, Abele and Kim 1989, Wicksten 1989, Rodríguez-Almaraz and Campos 1996, Mejía-Ortiz et al. 2001, Melo 2003, Valencia and Campos 2007, Almeida et al. 2008, Lara 2009, Mejía-Ortiz and Álvarez 2010, Lara and Wehrtmann 2011, Pileggi and Mantelatto 2012, Anger 2013, García-Guerrero et al. 2013, Rossi and Mantelatto 2013 Holthuis 1950, 1952, Wicksten 1989, March et al. 1998, Valencia and Campos 2007, Hein et al. 2011, Lara and Wehrtmann 2011, Anger 2013, García-Guerrero et al. 2013
Table 5.

Distributional and ecological comparison among each Macrobrachium species of sibling pair 5 and “6”.

Sibling 5 “Sibling 6”
M. tenellum M. acanthurus M. panamense M. amazonicum
American slope Pacific Atlantic Pacific Atlantic
Distribution Mexico (Baja California) to Peru USA (North Caroline) to Brazil (Rio Grande do Sul) Honduras to Peru South American river basins from Venezuela to Argentina
Habitat wide range of altitudes (more common in median courses of the rivers) wide range of altitudes (more common in higher elevations of the rivers)
Ecology require brackish water for reproduction (extended larval development with numerous and small eggs) require brackish water for reproduction (extended larval development with numerous and small eggs) inland (independent of salty water to reproduction) and coastal populations (dependent of salty water to reproduction) (distinct forms of extended larval development with numerous and small eggs)
Morphology similar and difficult to distinguish similar, and only few characters are useful features to separate both species
References Holthuis 1952, Choudhury 1970, Melo 2003, Hernández et al. 2007, Almeida et al. 2008, Lara and Wehrtmann 2011, Anger 2013, García-Guerrero et al. 2013. Holthuis 1952, Abele and Kim 1989, Melo 2003, Magalhães et al. 2005, Valencia and Campos 2007, Almeida et al. 2008, Lara 2009, Jara 2010, Vergamini et al. 2011, Anger 2013, Meireles et al. 2013

The time of divergence between both species of the Sibling 3 was approximately from 1.66 to 3.16 Mya for 16S gene, which supports the efficiency of the barrier in the separation of sibling species by mechanism of allopatric speciation. The morphologically close relation of the “olfersii complex” (see Villalobos 1969 for revision) was corroborated in the phylogeny, where Siblings 3 and 4 form sister groups with the same shape of the rostrum (Fig. 1B), as evidenced in previous molecular results (Rossi and Mantelatto 2013). The entity of the results obtained together with morphological and ecological similarities of M. olfersii and M. digueti suggest that both are sibling species, but the inclusion of other species from the M. olfersii complex in the analysis is necessary to confirm this proposition. Among the sibling species proposed by Holthuis (1952), only one pair (M. surinamicum × M. transandicum) was not analyzed in our study due the impossibility to obtain specimens of M. transandicum. In our phylogeny, M. surinamicum was included inside the clade of Macrobrachium olfersii complex (Villalobos 1969, Rossi and Mantelatto 2013) corresponding to a species with the rostrum almost straight, usually with more than 10 teeth in the upper margin (Fig. 1B).

For Siblings 2 and 4 the time of divergence between the species varied from 2.11 to 4.66 and 1.88 to 3.5 Mya for 16S gene, respectively. These data place them exactly in the range of the closure of the Isthmus, precluding the definition that the separation of the species may have been caused by this vicariant process. Pileggi and Mantelatto (2010) mentioned that M. americanum could be a synonymous of M. carcinus based on a single molecular 16S phylogeny. However, and as suggested by the authors, a more extensive sampling of M. americanum will be necessary to verify this proposition. Our results that include five specimens of M. carcinus and seven of M. americanum from distinct localities revealed that both species are sibling species.

In the same way, our data as well as the morphological and ecological similarities evidenced the close relationship between M. crenulatum and M. hancocki; however, the addition of data from more specimens and other species from the M. olfersii complex is necessary to confirm them as sibling species, i.e., sister taxa geographically separated (Rossi and Mantelatto, unpubl. data). Another unpredictable result was the close relation of M. michoacanus with M. hancocki (Fig. 1, Sibling 4). With both occurring on the Pacific side, this result may be interpreted as an indication that the relation of phylogenetically closely related congeners living on either side of the Isthmus must be older than the biogeographic barrier separating them (Anger 2013). New diversifications succeeding the closure of the Isthmus occurred at the same side, which can be demonstrated by the higher proximity between these sympatric species than M. hancocki and M. crenulatum, the hypothetical Sibling 4. However, analysis of additional material is necessary to verify the phylogenetic position of M. michoacanus. Following the other sibling species, relationship of the systematic position with the shape of the rostrum was maintained (Fig. 1B–C) supporting the high reliability of this morphological character.

Our results regarding M. amazonicum × M. panamense did not confirm a separate sibling group despite the close phylogenetic relation among these species. Our multigenic phylogenetic hypothesis (Fig. 1) indicates M. panamense as a sister group of the Sibling 5, and M. amazonicum as a sister species of this group (M. panamense + Sibling 5). Genetic divergence analyses of the “Sibling 6” pair (8.33 to 14.5 Mya for 16S genes) suggest that the time of their divergence predates the closure of the Isthmus, indicating that both did not share the same ancestral. In addition, the wide geographic distribution of M. amazonicum in the large South American river basins must be related to geological events driven by the rising Andes along the western portion of these basins (supposedly its native area of occurrence) (see Magalhães et al. 2005 for revision). Macrobrachium jelskii as an external clade of Sibling 5 and “Sibling 6” is in agreement with morphological similarities among these species, mainly of the shape of the rostrum (Fig. 1A), despite M. jelskii being the unique species of the group to present abbreviated larval development. The position of a more external group (M. potiuna, M. brasiliense, M. borellii) with abbreviated larval development in the phylogeny indicates that the ancestral species of this entire group possibly had a life cycle independent of salt water as suggested in previously studies (Murphy and Austin 2005, Pileggi and Mantelatto 2010). Macrobrachium amazonicum plays a key role in this puzzle since it presents inland and coastal populations (Vergamini et al. 2011, Meireles et al. 2013), suggesting that the species originated in freshwater environments and entered subsequently in estuarine habitats (Pereira and García 1995, Pileggi and Mantelatto 2010).

Phylogenetic analyses were based on two mitochondrial and one nuclear genes in order to provide a broad spectrum of inference and insights into the evolutionary history of Macrobrachium in the Americas. Although the mitochondrial markers may offer strong evidence for genus and species-level relationships, they have high mutation rates, which can cause increasing saturation when older splits are analyzed (Simon et al. 1994, Avise 2004). Therefore, analyses were carried out with sequences from conserved and variable genes to access phylogenetic information across a range of evolutionary time (Crandall et al. 2000). The genes concatenated analysis improve the diversity of evolutionary time, consequently revealed a more consistent phylogeny compared to previous morphological and molecular phylogenetics studies (Pereira 1997, Murphy and Austin 2005, Pileggi and Mantelatto 2010). The inclusion of the member of genus Cryphiops within Macrobrachium species was maintained in the phylogeny, and raises the question whether Macrobrachium is a monophyletic group (Pereira 1997, Murphy and Austin 2005, Pileggi and Mantelatto 2010, Carvalho et al. 2013, Rossi and Mantelatto 2013).

The results of our multidisciplinary approach suggest that species pairs 1-5 refer to siblings, in which each pair of species is difficult to distinguish using traditional morphological characters, although they are genetically distinct, closely related, and reproductively isolated (Mayr 1963, Steyskal 1972, Knowlton 1986). In contrast, our data did not validate “Sibling 6” by molecular analysis, although morphology, ecology, and geographic distribution patterns seem to suggest that they are sibling species (Holthuis 1952). Moreover, the speciation processes of the species of the pairs 2, 3, and 4 seem to have occurred after the rise of the isthmus barrier, probably in Pliocene and Pleistocene by the mechanism of allopatric speciation. However, the isolation of pairs 1 and 5 may have happened before the rise of the isthmus barrier, probably in Miocene by the mechanism of sympatric speciation.

An intriguing case refers to the occurrence of two species (M. hobbsi and M. olfersii) on both sides of the Central American land bridge (Anger 2013). The identification of these species may be incorrect or is related to the possibility of a passageway that has started with the opening of the Panama Canal, a scenario that has been already reported (Hildebrand 1939, Abele and Kim 1989). The possible expansion of the distribution range of Macrobrachium through the Panama Canal may occur, especially considering the dispersal potential of these amphidromous species (Bauer and Delahoussaye 2008, Bauer 2011, 2013). However the findings of the previous genetic study with M. olfersii revealed the absence of gene flow between Pacific and Atlantic populations. Moreover, M. digueti and specimens from Pacific considered as M. olfersii did not show divergence enough to split them, and the differences were within the range of interspecific values. Therefore, on the Pacific coast only M. digueti occurs naturally, which is considered, like M. olfersii, a sibling and cryptic species (Rossi and Mantelatto 2013).

This is the first phylogenetic study using molecular methods devoted entirely to the American transisthmian Macrobrachium sister species. Molecular markers confirmed that the Isthmus of Panama is an effective barrier contributing to the separation of sibling freshwater prawns species by the mechanism of allopatric speciation. However, some species seemed to have evolved before the closure of the Isthmus by the mechanism of sympatric speciation. Our phylogenetic analysis revealed consistent groups in most of the studied pairs endorsing the supposed sibling species. In contrast, the position of one pair (M. amazonicum × M. panamense) seems to be artificial once they did not share a recent common ancestor. The results presented here contribute to resolution of some doubts about the relationships of geminate American species. Our results support the conclusion that these sibling species are valid taxonomic entities, but not all transisthmian species are the closest living relatives with each other.

Acknowledgements

The present study is part of a long-term project to evaluate the taxonomy of freshwater/estuarine decapods in Brazil, and was supported by scientific grants provided to FLM by the Fundação de Amparo à Pesquisa do Estado de São Paulo - FAPESP (2002/08178-9; PhD to LGP 2005/50651-1; Biota 2010/50188-8; Coleções Científicas 2009/54931-0) and the Conselho Nacional de Desenvolvimento Científico e Tecnológico – CNPq - Brazil (472746/2004-9, 491490/2004-6, 490353/2007-0, 473050/2007-2, 471011/2011-8; 490314/2011, and 2504322/2012-5) and Consejo Nacional para Investigaciones Científicas y Tecnológicas CONICIT - Costa Rica (CII-001-08), during the development of the International Cooperative Project, which provided financial support to FLM, LGP, NR and ISW during the Brazil-Costa Rica visiting program, making possible the analysis of material and discussions. FLM also thanks CNPq for research grants (Research Scholarships PQ 301261/2004-0, 301359/2007-5 and 302748/2010-5), LGP is supported by post-doctoral fellowship (Proc. 02630/09-5) and NR is supported by PhD fellowship, both from Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - CAPES. We are grateful to the Department of Biology and Postgraduate Program in Comparative Biology of the FFCLRP/USP, and to many colleagues and friends (Alexandre Almeida, Cassiano Caluff, Darryl Felder, Edvanda Souza-Carvalho, Emerson Mossolin, Fernando Álvarez, Georgina Bond-Buckup, Harry Boos, José Luis Villalobos, Juan Bolaños, Luis Ernesto Arruda Bezerra, Marcos Tavares, Michael Türkay, Peter Schwendinger, Rodrigo Johnsson, Sérgio Althoff, Sérgio Bueno) for their help in collections, for making available some essential fresh specimens, for lending material from collections used in our research, and for critical discussion during the preparation of this manuscript. ISW would like to thank Luis Rólier Lara for his collaboration with collecting material in Costa Rica. Thanks are due to all members of LBSC for their assistance during the development of this study, and to anonymous reviewers and Ray Bauer for their valuable comments and suggestions.

References

  • Abele LG, Kim W (1989) The decapod crustaceans of the Panama Canal. Smithsonian Contributions to Zoology 482: 1–50. doi: 10.5479/si.00810282.482
  • Almeida AO, Coelho PA, Luz JR, Santos JTA, Ferraz NR (2008) Decapod crustaceans in fresh waters of southeastern Bahia, Brazil. Revista de Biología Tropical 56(3): 1225–1254.
  • Anger K (2013) Neotropical Macrobrachium (Caridea: Palaemonidae): on the biology, origin, and radiation of freshwater-invading shrimp. Journal of Crustacean Biology 33(2): 151–183. doi: 10.1163/1937240X-00002124
  • Avise JC (2004) Molecular Markers, Natural History and Evolution. 2ª ed. Sinauers Associates, Sunderland, 684 pp.
  • Bauer RT (2011) Amphidromy and migrations of freshwater shrimps. I. Costs, benefits, evolutionary origins, and an unusual case of amphidromy. New Frontiers in Crustacean Biology, 145–156. doi: 10.1163/ej.9789004174252.i-354.108
  • Bauer RT (2013) Amphidromy in shrimps: a life cycle between rivers and the sea. Latin American Journal of Aquatic Research 41(4): 633–650. doi: 10.3856/vol41-issue4-fulltext-2
  • Bauer RT, Delahoussaye J (2008) Life history migrations of the amphidromous river shrimp Macrobrachium ohione from a continental large river system. Journal of Crustacean Biology 28: 622–632. doi: 10.1651/08-2977.1
  • Bilton DT, Paula J, Bishop JDD (2002) Dispersal, genetic differentiation and speciation in estuarine organisms. Estuarine, Coastal and Shelf Science 55(6): 937–952. doi: 10.1006/ecss.2002.1037
  • Bowles DE, Aziz K, Knight CL (2000) Macrobrachium (Decapoda: Caridea: Palaemonidae) in the contiguous United States: a review of the species and an assessment of threats to their survival. Journal of Crustacean Biology 20(1): 158–171. doi: 10.1163/20021975-99990025
  • Bush GL (1994) Sympatric speciation in animals: new wine in old bottles. Trends in Ecology and Evolution 9(8): 285–288. doi: 10.1016/0169-5347(94)90031-0
  • Carvalho FL, Pileggi LG, Mantelatto FL (2013) Molecular data raise the possibility of cryptic species in the Brazilian endemic prawn Macrobrachium potiuna (Decapoda, Palaemonidae). Latin American Journal of Aquatic Research 41(4): 707–717. doi: 10.3856/vol41-issue4-fulltext-7
  • Choudhury PC (1970) Complete larval development of the palaemonid shrimp Macrobrachium acanthurus (Wiegmann, 1836), reared in the laboratory. Crustaceana 18: 113–132. doi: 10.1163/156854070X00743
  • Choudhury PC (1971) Complete larval development of the palaemonid shrimp Macrobrachium carcinus (L.), reared in the laboratory (Decapoda, Palaemonidae). Crustaceana 20(1): 51–69. doi: 10.1163/156854071X00526
  • Coates AG, Obando JA (1996) The geologic evolution of the Central American Isthmus. In: Jackson JBC, Budd AF, Coastes AG (Eds) Evolution and Environment in Tropical America. University of Chicago Press, Chicago, 21–56.
  • Coates AG, Jackson JBC, Collins LS, Cronin TM, Dowsett HJ, Bybell LM, Jung P, Obando JA (1992) Closure of the Isthmus of Panama: the near-shore marine record of Costa Rica and western Panama. Geological Society of America Bulletin 104: 814–828. doi: 10.1130/0016-7606(1992)104<0814:COTIOP>2.3.CO;2
  • Crandall KA, Harris DJ, Fetzner JW (2000) The monophyletic origin of freshwater crayfish estimated from nuclear and mitochondrial DNA sequences. Proceedings Biological Sciences/The Royal Society 267(1453): 1679–86. doi: 10.1098/rspb.2000.1195
  • Darriba D, Taboada GL, Doallo R, Posada D (2012) jModelTest 2: more models, new heuristics and parallel computing. Nature Methods 9(8): 772. doi: 10.1038/nmeth.2109
  • De Grave S, Fransen CHJM (2011) Carideorum catalogus: the recent species of the dendrobranchiate, stenopodidean, procarididean and caridean shrimps (Crustacea: Decapoda). Zoologische Mededelingen 85(9): 195–589.
  • Dugger DM, Dobkin S (1975) A contribution to knowledge of the larval development of Macrobrachium olfersii (Wiegmann, 1836) (Decapoda, Palaemonidae). Crustaceana 29(1): 1–30. doi: 10.1163/156854075X00018
  • Fitzpatrick BM, Fordyce JA, Gavrilets S (2008) What, if anything, is sympatric speciation? Journal of Evolutionary Biology 21(6): 1452–1459. doi: 10.1111/j.1420-9101.2008.01611.x
  • García-Guerrero MU, Becerril-Morales F, Vega-Villasante F, Espinosa-Chaurand LD (2013) Los langostinos del género Macrobrachium con importancia económica y pesquera en América Latina: conocimiento actual, rol ecológico y conservación. Latin American Journal of Aquatic Research 41(4): 651–675. doi: 10.3856/vol41-issue4-fulltext-3
  • Hedgpeth JW (1949) The North American species of Macrobrachium (river shrimp). Texas Journal of Science 1(3): 28–38.
  • Hein C, Pike A, Blanco J, Covich A, Scatena F, Hawkins C, Crowl T (2011) Effects of coupled natural and anthropogenic factors on the community structure of diadromous fish and shrimp species in tropical island streams. Freshwater Biology 56(5): 1002–1015. doi: 10.1111/j.1365-2427.2010.02537.x
  • Hernández L, Murugan G, Ruiz-Campos G, Maeda-Martínez A (2007) Freshwater shrimp of the genus Macrobrachium (Decapoda: Palaemonidae) from Baja California Peninsula, México. Journal of Crustacean Biology 27(2): 351–369. doi: 10.1651/S-2709.1
  • Hildebrand SF (1939) The Panama Canal as a passageway for fishes, with lists and remarks on the fishes and invertebrates. Zoologica 24(3): 15–45.
  • Holthuis LB (1950) Preliminary descriptions of twelve new species of Palemonid prawns from American waters (CrustaceaDecapoda). Proceedings Koninklijke Nederlandische Akademie van Wetenschapen 53: 93–99.
  • Holthuis LB (1952) A general revision of the Palaemonidae (Crustacea, Decapoda, Natantia) of the Americas II. The subfamily Palaemoninae. Occasional Paper, Allan Hancock Foundation Publications 12: 1–396.
  • Jara YG (2010) Morfometría y reproducción de tres especies langostinos de la vertiente del Pacífico de Costa Rica: Macrobrachium panamense, M. americanum y M. tenellum (Decapoda: Palaemonidae). Licenciatura Thesis, University of Costa Rica, San José, Costa Rica.
  • Jarman SN, Elliott NG, Nicol S, McMinn A (2011) Molecular phylogenetics of circumglobal Euphausia species (Euphausiacea: Crustacea). Canadian Journal of Fisheries and Aquatic Sciences 57(S3): 51–58. doi: 10.1139/f00-180
  • Jordan DS (1908) The law of the geminate species. The American Naturalist 42: 73–80. doi: 10.1086/278905
  • Keigwin LD (1978) Pliocene closing of the Isthmus of Panama, based on biostratigraphic evidence from nearby Pacific Ocean and Caribbean Sea cores. Geology 6: 630–634. doi: 10.1130/0091-7613(1978)6<630:PCOTIO>2.0.CO;2
  • Kimura M (1981) Estimation of evolutionary distances between homologous nucleotide sequences. Proceedings of the National Academy of Sciences 78: 454–458. doi: 10.1073/pnas.78.1.454
  • Kitaura J, Nishida M, Wada K (2002) Genetic and behavioral diversity in the Macrophthalmus japonicus species complex (Crustacea: Brachyura: Ocypodidae). Marine Biology 140(1): 1–8. doi: 10.1007/s002270100619
  • Knowlton N (1986) Cryptic and sibling species among the Decapod Crustacea. Journal of Crustacean Biology 6(3): 356–363. doi: 10.2307/1548175
  • Knowlton N, Weigt LA (1998) New dates and new rates for divergence across the Isthmus of Panama. Proceedings of the Royal Society B: Biological Sciences 265: 2257–2263. doi: 10.1098/rspb.1998.0568
  • Knowlton N, Weigt LA, Solórzano LA, Mills DK, Bermingham E (1993) Divergence in proteins, mitochondrial DNA, and reproductive compatibility across the Isthmus of Panama. Science 260: 1629–1632. doi: 10.1126/science.8503007
  • Lai JCY, Ng PKL, Davie PJF (2010) A revision of the Portunus pelagicus (Linnaeus, 1758) species complex (Crustacea: Brachyura, Portunidae), with the recognition of four species. Raffles Bulletin of Zoology 58(2): 199–237.
  • Lara LR (2009) Camarones dulceacuícolas (Decapoda: Palaemonidae) de la cuenca del río Grande de Térraba, Costa Rica. Proyeto Hidroeléctrico El Diquis, ICE, San José, 1–38.
  • Lara LR, Wehrtmann IS (2011) Diversity, abundance and distribution of river shrimps (Decapoda, Caridea) in the largest river basin of Costa Rica, Central America. New Frontiers in Crustacean Biology, 197–211.
  • Lessios HA (1998) The first stage of speciation as seen in organisms separated by the Isthmus of Panama. In: Howard DJ, S Berlocher (Eds) Endless Forms: Species and Speciation. Oxford University Press, Oxford, 186–201.
  • Lessios HA (2008) The great American schism: divergence of marine organisms after the rise of the Central American Isthmus. Annual Review of Ecology, Evolution, and Systematics 39: 63–91. doi: 10.1146/annurev.ecolsys.38.091206.095815
  • Magalhães C, Bueno SLS, Bond-Buckup G, Valenti WC, Silva HLM, Kiyohara F, Mossolin EC, Rocha SS (2005) Exotic species of freshwater decapod crustaceans in the state of São Paulo, Brazil: records and possible causes of their introduction. Biodiversity and Conservation 14(8): 1929–1945. doi: 10.1007/s10531-004-2123-8
  • Malay MCD, Paulay G (2009) Peripatric speciation drives diversification and distributional pattern of reef hermit crabs (Decapoda: Diogenidae: Calcinus). Evolution 64(3): 634–662. doi: 10.1111/j.1558-5646.2009.00848.x
  • Mantelatto FL, Pileggi LG, Miranda I, Wehrtmann IS (2011) Does Petrolisthes armatus (Anomura, Porcellanidae) form a species complex or are we dealing with just one widely distributed species? Zoological Studies 50(3): 372–384.
  • March J, Benstead P, Pringle M, Scatena M (1998) Migratory drift of larval freshwater shrimps in two tropical streams, Puerto Rico. Freshwater Biology 40: 261–273. doi: 10.1046/j.1365-2427.1998.00352.x
  • Marko PB (2002) Fossil calibration of molecular clocks and the divergence times of geminate species pairs separated by the Isthmus of Panama. Molecular Biology and Evolution 19(11): 2005–2021. doi: 10.1093/oxfordjournals.molbev.a004024
  • Meireles AL, Valenti WC, Mantelatto FL (2013) Reproductive variability of the Amazon River prawn, Macrobrachium amazonicum (Caridea, Palaemonidae): influence of life cycle on egg production. Latin American Journal of Aquatic Research 41(4): 718–731. doi: 10.3856/vol41-issue4-fulltext-8
  • Mejía-Ortiz LM, Alvarez F (2010) Seasonal patterns in the distribution of three species of freshwater shrimp, Macrobrachium spp., along an altitudinal river gradient. Crustaceana 83(4): 385–397. doi: 10.1163/001121610X489368
  • Mejía-Ortíz LM, Alvarez F, Román R, Viccon-Pale JA (2001) Fecundity and distribution of freshwater prawns of the genus Macrobrachium in the Huitzilapan river, Veracruz, Mexico. Crustaceana 74(1): 69–77. doi: 10.1163/156854001505442
  • Melo GAS (2003) Famílias Atyidae, Palaemonidae e Sergestidae. In: Melo GAS (Ed) Manual de Identificação dos Crustáceos Decápodos de Água Doce Brasileiros. Editora Loyola, São Paulo, 289–415.
  • Miura O, Torching ME, Bermingham E (2010) Molecular phylogenetics reveals differential divergence of coastal snails separated by the Isthmus of Panama. Molecular Phylogenetics and Evolution 56(1): 40–48. doi: 10.1016/j.ympev.2010.04.012
  • Monaco G (1975) Laboratory rearing of the larvae of the palaemonid shrimp Macrobrachium americanum (Bate). Aquaculture 6: 369–375. doi: 10.1016/0044-8486(75)90115-5
  • Moraes-Riodades PMC, Valenti WC (2004) Morphotypes in male Amazon river prawns, Macrobrachium amazonicum. Aquaculture 236(1–4): 297–307. doi: 10.1016/j.aquaculture.2004.02.015
  • Morrison CL, Ríos R, Duffy JE (2004) Phylogenetic evidence for an ancient rapid radiation of Caribbean sponge-dwelling snapping shrimps (Synalpheus). Molecular Phylogenetics and Evolution 30: 563–581 doi: 10.1016/S1055-7903(03)00252-5
  • Murphy N, Austin C (2005) Phylogenetic relationships of the globally distributed freshwater prawn genus Macrobrachium (Crustacea: Decapoda: Palaemonidae): biogeography, taxonomy, and the convergent evolution of abbreviated larval development. Zoologica Scripta 34(2): 187–197. doi: 10.1111/j.1463-6409.2005.00185.x
  • Negri M, Pileggi LG, Mantelatto FL (2012) Molecular barcode and morphological analyses reveal the taxonomic and biogeographical status of the striped-legged hermit crab species Clibanarius sclopetarius (Herbst, 1796) and Clibanarius vittatus (Bosc, 1802) (Decapoda: Diogenidae). Invertebrate Systematics 26: 561–571. doi: 10.1071/IS12020
  • Palumbi SR, Benzie J (1991) Large mitochondrial DNA differences between morphologically similar penaeid shrimp. Molecular Marine Biology and Biotechnology 1(1): 27–34.
  • Palumbi SR, Martin A, Romano S, McMillan WO, Stice L, Grabowski G (1991) The Simple Fool’s Guide to PCR. Department of Zoology and Kewalo Marine Laboratory, University of Hawaii, Honolulu.
  • Pereira G (1997) A cladistic analysis of the freshwater shrimps of the family Palaemonidae (Crustacea, Decapoda, Caridae). Acta Biologica Venezuelica 17: 1–69.
  • Pereira G, García DJV (1995) Larval development of Macrobrachium reyesi Pereira (Decapoda: Palaemonidae), with a discussion on the origin of abbreviated development in palaemonids. Journal of Crustacean Biology 15: 117–133. doi: 10.2307/1549016
  • Pileggi LG, Mantelatto FL (2010) Molecular phylogeny of the freshwater prawn genus Macrobrachium (Decapoda, Palaemonidae), with emphasis on the relationships among selected American species. Invertebrate Systematics 24(2): 194–208. doi: 10.1071/IS09043
  • Pileggi LG, Mantelatto FL (2012) Taxonomic revision of doubtful Brazilian freshwater shrimp species of genus Macrobrachium (Decapoda, Palaemonidae). Iheringia, Série Zoologia 102(4): 426–437. doi: 10.1590/S0073-47212012005000012
  • Pileggi LG, Magalhães C, Bond-Buckup G, Mantelatto FL (2013) New records and extension of the known distribution of some freshwater shrimps in Brazil. Revista Mexicana de Biodiversidad 84: 563–574. doi: 10.7550/rmb.30504
  • Porter ML, Pérez-Losada M, Crandall KA (2005) Model-based multi-locus estimation of decapod phylogeny and divergence times. Molecular Phylogenetics and Evolution 37: 355–369. doi: 10.1016/j.ympev.2005.06.021
  • Posada D, Buckley T (2004) Model selection and model averaging in phylogenetics: advantages of Akaike information criterion and Bayesian approaches over likelihood ratio tests. Systematic Biology 53: 793–808.
  • Rocha SS, Bueno SLS (2004) Crustáceos decápodes de água doce com ocorrência no Vale do Ribeira de Iguape e rios costeiros adjacentes, São Paulo, Brasil. Revista Brasileira de Zoologia 21(4): 1001–1010. doi: 10.1590/S0101-81752004000400038
  • Rodríguez-Almaraz GA, Campos E (1996) New locality records of freshwater decapods from México (Crustacea: Atyidae, Cambaridae and Palaemonidae). Proceedings of the Biological Society of Washington 109: 34–38.
  • Rossi N, Mantelatto FL (2013) Molecular analysis of the freshwater prawn Macrobrachium olfersii (Decapoda, Palaemonidae) supports the existence of a single species throughout its distribution. PLoS ONE 8(1): e54698. doi: 10.1371/journal.pone.0054698
  • Schubart CD, Neigel JE, Felder DL (2000) Use of the mitochondrial 16S rRNA gene for phylogenetic and population studies of Crustacea. Crustacean Issues 12: 817–830.
  • Schubart CD, Cuesta JA, Rodríguez A (2001b) Molecular phylogeny of the crab genus Brachynotus (Brachyura: Varunidae) based on the 16S rRNA gene. Hydrobiologia 449(1): 41–46. doi: 10.1007/978-94-017-0645-2_3
  • Schubart CD, Conde JE, Carmona-Suárez C, Robles R, Felder DL (2001a) Lack of divergence between 16S mtDNA sequences of the swimming crabs Callinectes bocourti and C. maracaiboensis (Brachyura: Portunidae) from Venezuela. Fishery Bulletin 99(3): 475–481.
  • Simon C, Beckenbach A, Crespi B, Lui H, Flook P (1994) Evolution, weighting and phylogenetic utility of mitochondrial gene sequences and compilation of conserved polymerase chain reactions primers. Annals of the Entomological Society of America 87: 651–468.
  • Stanhope MJ, Leighton BJ, Hartwick B (1992) Polygenic control of habitat preference and its possible role in sympatric population subdivision in an estuarine crustacean. Heredity 69: 279–288. doi: 10.1038/hdy.1992.126
  • Steeves TE, Anderson DJ, Friesen VL (2005) The Isthmus of Panama: a major physical barrier to gene flow in a highly mobile pantropical seabird. Journal of Evolutionary Biology 18: 1000–1008. doi: 10.1111/j.1420-9101.2005.00906.x
  • Steyskal GC (1972) The meaning of the term “Sibling Species”. Systematic Zoology 21(4): 446. doi: 10.2307/2412441
  • Stillman JH, Reeb CA (2001) Molecular phylogeny of eastern pacific porcelain crabs, genera Petrolisthes and Pachycheles, based on the mtDNA 16S rDNA sequence: phylogeographic and systematic implications. Molecular Phylogenetics and Evolution 19(2): 236–245. doi: 10.1006/mpev.2001.0924
  • Swofford DL (2003) PAUP. Phylogenetic Analysis Using Parsimony (and Other Methods). Version 4. Sinauer Associates, Sunderland, Massachusetts.
  • 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(10): 2731–2739. doi: 10.1093/molbev/msr121
  • Thompson JD, Higgins DG, Gibson TJ (1994) CLUSTALW: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting specific gap penalties and weight matrix choice. Nucleic Acids Research 22: 4673–4680 doi: 10.1093/nar/22.22.4673
  • Torati LS, Mantelatto FL (2012) Ontogenetic and evolutionary change of external morphology of the neotropical shrimp Potimirim (Holthuis, 1954) explained by a molecular phylogeny of the genus. Journal of Crustacean Biology 32(4): 625–640. doi: 10.1163/193724012X635322
  • Valencia DM, Campos MR (2007) Freshwater prawns of the genus Macrobrachium Bate, 1868 (Crustacea: Decapoda: Palaemonidae) of Colombia. Zootaxa 1456: 1–44.
  • Vergamini FG, Pileggi LG, Mantelatto FL (2011) Genetic variability of the Amazon River prawn Macrobrachium amazonicum (Decapoda, Caridea, Palaemonidae). Contributions to Zoology 80(1): 67–83.
  • Villalobos FA (1969) Problemas de especiación en América de un grupo de Palaemonidae del genero Macrobrachium. In: Mistakidis MN (Ed.) FAO Fisheries and Aquaculture Report 57(3): 589–1165.
  • Villalobos-Hiriart JL, Álvarez F, Hernández C, De la Lanza-Espino G, González-Mora ID (2010) Crustáceos decápodos de las cuencas Copalita, Zimatán y Coyula, en Oaxaca, México. Revista Mexicana de Biodiversidad 81: 99–111.
  • Wehrtmann IS, Albornoz L (2002) Evidence of different reproductive traits in the transisthmian sister species, Alpheus saxidomus and A. simus (Decapoda, Caridea, Alpheidae): description of the first postembryonic stage. Marine Biology 140: 605–612.
  • Whiting MF, Carpenter JC, Wheeler QD, Wheeler WC (1997) The trepsiptera problem: phylogeny of the holometabolous insect orders inferred from 18S and 28S ribosomal DNA sequences and morphology. Systematics Biology 46: 1–68.
  • Wicksten MK (1989) A key to the palaemonid shrimp of the Eastern Pacific Region. Bulletin Southern California Academy of Science 88: 11–20.
login to comment