Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Molecular Analysis of the Freshwater Prawn Macrobrachium olfersii (Decapoda, Palaemonidae) Supports the Existence of a Single Species throughout Its Distribution

  • Natália Rossi ,

    nataliarossi.bio@gmail.com (NR); flmantel@usp.br (FM)

    Affiliation Laboratory of Bioecology and Crustacean Systematics (LBSC), Program in Comparative Biology, Department of Biology, Faculty of Philosophy, Science and Letters of Ribeirão Preto (FFCLRP), University of São Paulo (USP), Ribeirão Preto, São Paulo, Brazil

  • Fernando Luis Mantelatto

    nataliarossi.bio@gmail.com (NR); flmantel@usp.br (FM)

    Affiliation Laboratory of Bioecology and Crustacean Systematics (LBSC), Program in Comparative Biology, Department of Biology, Faculty of Philosophy, Science and Letters of Ribeirão Preto (FFCLRP), University of São Paulo (USP), Ribeirão Preto, São Paulo, Brazil

Abstract

Macrobrachium olfersii is an amphidromous freshwater prawn, widespread along the eastern coasts of the Americas. This species shows great morphological modifications during ontogenesis, and several studies have verified the existence of a wide intraspecific variation. Because of this condition, the species is often misidentified, and several synonyms have been documented. To elucidate these aspects, individuals of M. olfersii from different populations along its range of distribution were investigated. The taxonomic limit was established, and the degree of genetic variability of this species was described. We extracted DNA from 53 specimens of M. olfersii, M. americanum, M. digueti and M. faustinum, which resulted in 84 new sequences (22 of 16S mtDNA, 45 of Cythocrome Oxidase I (COI) mtDNA, and 17 of Histone (H3) nDNA). Sequences of three genes (single and concatenated) from these species were used in the Maximum Likelihood and Bayesian Inference phylogenetic analyses and COI sequences from M. olfersii were used in population analysis. The genetic variation was evaluated through the alignment of 554 bp from the 16S, 638 bp from the COI, and 338 bp from the H3. The rates of genetic divergence among populations were lower at the intraspecific level. This was confirmed by the haplotype net, which showed a continuous gene flow among populations. Although a wide distribution and high morphological intraspecific variation often suggest the existence of more than one species, genetic similarity of Caribbean and Brazilian populations of M. olfersii supported them as a single species.

Introduction

Prawns of the genus Macrobrachium Spence Bate, 1861 are important members of freshwater and estuarine systems [1], [2]. The diversity of species is tremendous, with more than 240 species recognized worldwide [3], [4]. Some of them need estuarine and freshwater to complete their life cycle, which implicates in recurring migrations between both environments. The prawns that have this behavior are called as amphidromous and they show many larval and reproductive peculiarities [5]. In some amphidromous freshwater shrimps, the females live in freshwater and migrate downstream, near to estuaries, where the hatching occurs. After larval development in salt water, the juveniles migrate up to the freshwater [6]. Macrobrachium olfersii (Wiegmann, 1836) is one these species, with larval stages requiring high salinity during development [7], [8]. Therefore, it is finding along of coastland and its geographic distribution covers almost all the eastern coasts of the Americas, from the southern United States, where the prawns were introduced [9], [10], [11], to Mexico, Guatemala, Costa Rica, Panama, and Venezuela to southern Brazil [12], [9], [11], [13], [14], [15].

Species that have a wide distribution, in heterogeneous or geographically isolated environments can have a phenotype variation, because they are prone to show plastic responses to different environmental influences. Further these plasticity environment-dependent, the phenotypic variations also can stem from genetic or behavior differences between individuals, from ontogenetic developmental or combining of all these factors [16]. On the other hand, morphological characters may often be undergoing convergent evolution as they are under similar selective pressure [17]. The species of the genus Macrobrachium have high intraspecific variation and a single species may have genetic diversity and structured populations [18].

Macrobrachium olfersii is in this context, it has a wide distribution and a great morphological variation during its ontogenesis [19], [20], [21], [22], including the possibility of morphotypes within the species. This fact have already observed in other congeneric species such as Macrobrachium rosenbergii (de Man, 1879), M. amazonicum (Heller, 1862) and M. grandimanus [23], [24], [25]. The presence of plastic characters in the genus Macrobrachium makes the accurate determination of species more difficult and problematic [10], [26].

Commonly, Macrobrachium olfersii is confused with species that occur preferentially in Central America such as M. faustinum (Saussure, 1857), M. crenulatum Holthuis, 1950, M. digueti (Bouvier, 1895), M. hancocki Holthuis, 1950, and M. acanthochirus Villalobos, 1967. Because of their close morphological similarities, these species were designated as a possible species complex by Villalobos [10]. Recently, Macrobrachium birai Lobão, Melo & Fernandez, 1986 and Macrobrachium holthuisi Genofre & Lobão, 1978 were considered junior synonyms of M. olfersii [27], [28].

Together, this information illustrates that sometimes morphological analysis alone is not sufficient to resolve the diversity of species complexes. The addition of molecular data has proven very useful to elucidate the taxonomic relationships in morphologically variable groups of freshwater prawns [29], [2], [27], [18], [30], [31]. Molecular markers can be useful in delimiting boundaries between lineages and/or species and can help in the interpretation of biodiversity patterns [32], [33], [27], [18], [31]. The mitochondrial markers, such as 16S and Cytochrome Oxidase I (COI), have high mutation rate, which makes it very useful at intraspecific levels, but causes increasing saturation when older splits are analyzed. Therefore, the combination of mtDNA with more conserved nuclear markers, such as Histone 3 (H3), may be used to support phylogenetic conclusions [34], [35]. The H3 is considered a small gene with about 350 bp, which are very conservative and relative ease to amplify. When we combined nuclear with mitochondrial genes, we provide a broad spectrum of inference and great insights into the evolutionary history of Macrobrachium.

Considering that Macrobrachium inhabits heterogeneous environments and shows morphological dissimilarity and genetic variability within populations [18], this study examined individuals of M. olfersii from different populations along its distribution, in order to establish the taxonomic boundaries and to describe the degree of genetic variability.

Materials and Methods

Sample Collection

Some specimens of M. olfersii were obtained from field collections in different locations, under license from the appropriate authorities (Table 1). The collections of species conducted in this study complied with current applicable state and federal laws of Brazil (DIFAP/IBAMA 126/05; permanent license to FLM for collection of Zoological Material No. 11777-1 MMA/IBAMA/SISBIO). No material were obtained from national park or other protected area of land and we confirm that the field studies did not involve endangered or protected species.

thumbnail
Table 1. Sequences of Macrobrachium and outgroup species used this study.

https://doi.org/10.1371/journal.pone.0054698.t001

We have collected about six individuals per site from coastal drainage of Costa Rica, Panama and different places from Brazil (States of Bahia, Espírito Santo, Rio de Janeiro, São Paulo, Paraná and Santa Catarina) (Table 1), by sieving amongst marginal vegetation and under the rocky bottom of rivers and streams. After sampling, individuals were separated, immediately preserved in ethanol (80%) and deposited in Crustacean Collection of the Biology Department (CCDB) of the Faculty of Philosophy, Sciences and Letters of Ribeirão Preto (FFCLRP) at the University of São Paulo (USP) (Permanent license for Crustacean Collection No. 071/2012/SECEX/CGEN).

Additional material was also obtained via donation or loan from the following museums and crustacean collections: Museum of Zoology, University of São Paulo, São Paulo, Brazil (MZUSP); Coleccíon Nacional de Crustáceos de La Universidad Nacional Autonoma de Mexico, Mexico City, Mexico (CNCR); Instituto Venezolano de Investigações Científicas, Venezuela (IVIC); Museum of Zoology, University of Costa Rica (MZUCR), Rijksmuseum Van Natuurlijke Historie, Leiden, Holland (RMNH) and (Table 1).

By these means of sampling (field collection, donation or loan of species from museums and crustacean collection), we were able to include in our analysis specimens from almost the entire range of distribution of M. olfersii, in order to have the most robust possible data set. Unfortunately, species from United States and Guatemala were not sampled. Nevertheless, our sampling effort was enough to analyze the variability of M. olfersii due to wide distribution this species.

Macrobrachium faustinum and M. digueti, which belong to the supposed M. olfersii complex [10], were included in our analysis because they are closest species to M. olfersii. Macrobrachium americanum Spence Bate, 1868, from Pacific coast of South and Central America, Macrobrachium lar (Fabricius, 1798) from Indo-Pacific and Cryphiops caementarius (Molina, 1782) from Pacific coast of South America were added. Previous phylogenetic analyses showed that they are closely related to species of the M. olfersii complex [36], [32], [37], [27] (Table 1).

Macrobrachium americanum and M. lar have similar life histories with extended larval development, and similar and diverse geographic distributions. Although C. caementarius belong to a different genus, it has been positioned among species of Macrobrachium [27]. This questionable phylogenetic position have shown a closely relationship with M. olfersii complex [27]. Like the sampling methods of the M. olfersii specimens, the exemplars of the M. faustinum, M. digueti and M. americanum were obtained by field collecting from Costa Rica and Panama, others exemplars from Mexico, Jamaica and Venezuela were loaned of the different museums and crustacean collections (described at Table 1).

Species Identification

Considering the previously reported taxonomic doubts of Macrobrachium olfersii complex of species, a detailed comparative study of external morphology of the group (M. acanthochirus, M. crenulatum, M. digueti, M. faustinum, M. hancocki and M. olfersii) was also conducted, and a key for identification of male adults of the M. olfersii species-complex will be proposed based on morphological analysis of greater pereiopod (data not shown here). In order to better understand the interspecies differences in morphological characters and provide us a robust support during specimen identification and molecular analysis (Rossi & Mantelatto, in preparation). The identifications were based on the diagnostic morphological traits of M. olfersii and related species, in accordance with the literature [9], [10], [38], [15].

Some characteristics that differentiate the species of M. olfersii complex are based mainly on shape, ornament and morphometric ratio of the articles of the second pereiopod, such as the number and distribution of tooth on the upper and cutting edge of the fingers, the format of the lower margin of the palm and the ratios obtained from the carpus length/merus length. On second pereiopod of the M. olfersii, there are tooth on all upper margin and the lower margin is convex. The carpus is equal or smaller than merus [9,10; Rossi & Mantelatto, in preparation]. However, these small variations can often be different stages of the development or stem as phenotypic plasticity [39], [10].

DNA Extraction, Amplification and Sequencing

Most sequences obtained in this study were generated from our own extractions for this project. These sequences were deposited to Genbank under the accession numbers listed in Table 1. Seven additional comparative sequences from Macrobrachium lar, M. olfersii and Cryphiops caementarius were retrieved from GenBank. Genetic vouchers were deposited in appropriate scientific zoological collections. DNA was extracted from abdominal or pereiopod muscle tissue from 53 Macrobrachium specimens from different localities (Table 1 and Fig. 1). When possible, the sequences were obtained from multiple representatives from each collection site.

thumbnail
Figure 1. Locations of Macrobrachium olfersii samples used to extract DNA.

1: CR-AT; 2: PN-AT; 3: VZ-NW; 4: VZ-IM; 5: BR-BA; 6: BR-ES; 7: BR-RJ; 8: BR-SPn; 9: BR-SPs; 10: BR-PR; 11: BR-SC. Abbreviations: see Table 1.

https://doi.org/10.1371/journal.pone.0054698.g001

DNA extraction and amplification procedures followed Mantelatto et al. [40], [41], [42] and Pileggi and Mantelatto [27]. Extracted tissues was macerated and incubated for 12–24 hs in 600 ml of lysis buffer at 65°C. Protein was separated with the addition of 200 ml of ammonium acetate (7.5 M), followed by centrifugation. The addition of 600 ml of cold (−20°C) absolute isopropanol was used for DNA precipitation, also followed by centrifugation. The resulting pellet was then washed with 70% ethanol, dried in a lyophilizer and resuspended in 20 ml of TE buffer.

DNA was amplified by polymerase chain reaction (PCR) using previously tested primers [43], [44]: H3ar (ATA TCC TTR GGC ATR ATR GTG AC) and H3af (ATG GCT CGT ACC AAG CAG ACV GC) for the Histone (H3) gene [45], H2 (AGA TAG AAA CCA ACC TGG) and L2 (TGC CTG TTT ATC AAA AAC AT) for the 16S gene [46], and COIa (AGT ATA AGC GTC TGG GTA GTC) and COIf (CCT GCA GGA GGA GGA GAC CC) for the Cytochrome Oxidase I (COI) gene [47].

Reactions were performed in 25 µl volumes containing 6.5 µl of distilled water, 3 µl of 10X PCR buffer II, 3 µl of MgCl2 (25 mM), 5 µl of betaine (5 M), 1 µl of each primer (10 mM), 4 µl of dNTP (10 mM), 0.5 µl of AmpliTaq DNA polymerase and1 µl of DNA. Thermal cycling was performed as follows for COI: initial denaturation for 2 min at 94°C, followed by 30 cycles of 30 s at 94°C, 30 s at 50°C, and 60 s at 72°C, with a final extension of 6 min at 72°C. Thermal cycling for 16S and H3∶5 min at 98°C, followed by 40 cycles of 45 s at 98°C, 45 s at 48°C, and 45 s at 72°C, with a final extension of 8 min at 72°C. The results of PCRs were looked at electrophoresis with agarose gel (1%).

PCR products were purified using the SureClean Plus kit, and sequenced with the ABI Big Dye® Terminator Mix in an ABI Prism 3100 Genetic Analyzer® following Applied Biosystems protocols. All sequences were confirmed by sequencing both strands. The consensus sequence for the two strands was obtained using BioEdit Version 7.0.7.1 [48]. Sequences were aligned in Clustal W with interface in BioEdit, with the following parameters: gap opening 10 and gap extending 0.2 [49]. Primer and indeterminate regions in beginning of the sequences were cut. The absence of stop codons in COI sequences was checked using the software BioEdit and the invertebrate codon table implemented in Mega 4 [50] in order to confirm the nonexistence of pseudogenes [51]. Apart from that, the consensus sequences were blast on Genbank and compared with our previous sequences.

Phylogenetic Analyses

All analyses were based on a partial fragment of the 16S mtDNA, COI mtDNA and H3 nDNA genes. The phylogenetic reconstructions were built by the Maximum Likelihood (ML) method in PAUP version 4.0 beta 10 [52]. The appropriate model of evolution was previously selected under the Akaike Information Criterion (AIC) [53] obtained from the jModelTest program [54]. Heuristic searches were used for ML analyses with 100,000 replicates of random sequence additions, and nonparametric bootstrapping consisted of 100 replications [55] with 10 random sequence additions in PAUP. Only values >50% were reported. In order to estimate intra- and interspecific divergence rates, genetic distances were also calculated in PAUP using the uncorrected p distance.

Moreover, phylogenetic hypothesis were also generated by Bayesian Inference - BI in the MrBayes 3.1 program [56] for each gene data and for concatenated genes (three genes and mitochondrial). Bayesian analysis was configured to use the following parameters: sampling frequency 500, four-chain heating (three heated and one cold), and the value of Stop heating chains less than 0.01 after at least 2.5 millions of generations. Subsequently, data were collected from stationary phase and chain the initial states discarded (burning = 15%). The levels of branch support were obtained by the method of posterior probability. Trees generated from both analyzes were saved and edited by Figtree program v.1.3.1 [57].

The haplotype number of the COI sequences from at least three individuals of each studied population of the M. olfersii was calculated in DnaSP Version 4.10.9 [58]. The haplotype and nucleotide diversities were calculated for each population using Arlequin Version 3.1 [59]. Haplotype networks were constructed by the statistical parsimony method in TCS (Version 1.21) [60] and by the Median-Joining method in Network [61], with data preparation in DnaSP. Series of analyses of molecular variance (AMOVA) [62] were computed in Arlequin to examine the distribution of genetic variation. Analyses were run based on haplotype frequencies with no hierarchical structure (all populations in a single group) and with a subdivision between Caribbean and Brazilian populations of M. olfersii. The significance was tested using a nonparametric permutation procedure [62], incorporating 10,000 permutations.

Results

In total, 22 sequences with 564 bp of 16S mtDNA, 45 sequences with 638 bp of COI without pseudogenes, and 17 sequences with 338 bp of H3 were generated. The frequencies of bases for each gene and the optimal models selected under AIC to use on Maximum Likelihood analyses (Table 2). The utilized model on Bayesian analysis was GTR for all data set.

thumbnail
Table 2. The frequencies of bases of the analyzing genes and the best-fit models selected under Akaike Information Criterion (AIC) on jModeltest for Maximum Likelihood (ML) analysis.

https://doi.org/10.1371/journal.pone.0054698.t002

Similar topologies were achieved on ML and BI analyses for each gene datasets. Therefore, the trees obtained by BI with posterior probability were revealed and the bootstrap obtained by ML were added (Figs. 2, 3, 4, 5). Disregarding the partial sequence of 16S from M. olfersii that was retrieved from GenBank (ID: AY377849), phylogenetic analyses (Figs. 2, 3 and 5) show an evident separation clade from all studied species, but in nuclear, H3 (Fig. 4). Although, the species were not separated in all cases, on BI analysis based on three concatenated genes, we can see a better resolution of the branches and the same topology was obtained on mitochondrial concatenated genes (Fig. 5). Moreover, M. olfersii and M. digueti formed a sister clade with M. faustinum in the phylogenetic tree of 16S, COI and concatenated genes.

thumbnail
Figure 2. Phylogenetic tree for populations of Macrobrachium olfersii, based on Bayesian Inference analysis of 16S data sets.

Abbreviations and code: see Table 1. Numbers on right: posterior probabilities; Numbers on left: bootstrap obtained on Maximum Likelihood analysis.

https://doi.org/10.1371/journal.pone.0054698.g002

thumbnail
Figure 3. Phylogenetic tree for populations of Macrobrachium olfersii, based on Bayesian Inference analysis of COI data sets.

Abbreviations and code: see Table 1. Numbers on right: posterior probabilities; Numbers on left: bootstrap obtained on Maximum Likelihood analysis.

https://doi.org/10.1371/journal.pone.0054698.g003

thumbnail
Figure 4. Phylogenetic tree for populations of Macrobrachium olfersii, based on Bayesian Inference analysis of H3 data sets.

Abbreviations and code: see Table 1. Numbers on right: posterior probabilities; Numbers on left: bootstrap obtained on Maximum Likelihood analysis.

https://doi.org/10.1371/journal.pone.0054698.g004

thumbnail
Figure 5. Phylogenetic tree for populations of Macrobrachium olfersii, based on Bayesian Inference analysis of three concatenated genes (16S, COI and H3).

Abbreviations and code: see Table 1. Numbers: posterior probabilities; Asterisk numbers: posterior probabilities obtained by Bayesian Inference analysis of mitochondrial genes (16S and COI concatenated).

https://doi.org/10.1371/journal.pone.0054698.g005

In general, distance analyses showed the percentage of intraspecific was lower than interspecific variation. M. olfersii data ranged from 0.00 to 0.18% for 16S, 0.00 to 0.95% for COI, and 0.00 to 0.27% for H3. The obtained values between M. olfersii and close species were: M. digueti, 1.25% for 16S, ranging from 10.18 to 10.72% for COI and from 0.00 to 0.27% for H3; M. faustinum, 3.84% for 16S, ranging from 12.51 to 12.70% for COI and from 0.27 to 0.54% for H3. Among other Macrobrachium species, the result was higher than before, e.g., M. americanum was 7.70% for 16S, ranging from 17.42 to 17.81% for COI and ranging from 5.92 to 6.26% for H3.

Based on a COI fragment of unambiguous sequence, 28 haplotypes (H) were recognized, which showed a total haplotype diversity of 0.94. Among the haplotypes, 25 (89.28%) represented single individuals, and 3 (10.72%) were polymorphic. The frequencies of haplotypes in different localities were heterogeneous (Table 3). The first haplotype (H1) was shared among three individuals from Paraná (BR-PR), two individuals from northern São Paulo (BR-SPn), one individual from Espírito Santo (BR-ES), one individual from Santa Catarina (BR-SC), one individual from Costa Rica (CR-AT), and one individual from Panama (PN-AT). The third haplotype (H3) was shared between one individual from BR-SPn and one from Costa Rica (CR-AT). The fourth (H4) was shared among one individual from BR-SC, one from BR-SPs, one individual from Bahia (BR-BA), and another from BR-RJ.

thumbnail
Table 3. Distribution of haplotypes detected in Macrobrachium olfersii from different localities.

https://doi.org/10.1371/journal.pone.0054698.t003

Network haplotypes constructed based on the statistical parsimony method (data not shown) were equal to network haplotypes constructed based on the Median-Joining method (Fig. 6). Analysis of molecular variance (AMOVA) indicated that specimens within the M. olfersii population have the highest percentage of variation (98.25% without hierarchical structure and 98.18% with hierarchical structure), whereas the variation among populations was low (1.75% without hierarchical structure and 1.69% with hierarchical structure). When populations were structured according to Caribbean and Brazilian populations, the variations among groups were very low (0.13%). However the values obtained by AMOVA based on haplotype frequencies with and without hierarchical structure were not significant (Table 4).

thumbnail
Figure 6. Haplotype network based on Median-Joining analysis, indicating the distribution of each haplotype found in Macrobrachium olfersii.

The haplotype identification is below each circle. Each small trace represents a mutational step. Lines between circles indicate single-substitution differences between haplotypes (the small symbol indicates one missing haplotype inferred for two sequences that differed by two substitutions), and are not proportional to the genetic distance between haplotypes. Pattern coding indicates the origin and the frequency of the shared haplotype, indicated in the legend at right. Abbreviations: see Table 1.

https://doi.org/10.1371/journal.pone.0054698.g006

thumbnail
Table 4. Analysis of molecular variance in Macrobrachium olfersii.

https://doi.org/10.1371/journal.pone.0054698.t004

Discussion

Our molecular results confirm the taxonomic status of M. olfersii as a valid species by analysis of individuals from Caribbean and Brazilian regions and a close phylogenetic relationship with other related species. Despite the absence of all species within the complex, we chose the closest species (M. digueti and M. faustinum), which really could be synonymous from M. olfersii. As far as we know, after about 43 years since the proposal of a possible morphological species complex, this is the most thorough study using the molecular phylogeny approach to elucidate the status of this species. Different analyses were computed and compared. This methodology had given consistent results, because we can see similar topologies and inferences.

Maximum likelihood (ML) has been considered a good reconstruction method in studies with Decapods [63], [27], [18], [64], [31]. These results were compared with obtained by Bayesian Inference (BI) that has also been efficient [65], [63], [66]. Similar topologies were achieved, although BI showed better resolution of the branches with high posterior probabilities.

We opted by a phylogenetic approach, firstly because of we followed the Phylogenetic Species Concepts sensu Mishler & Theriot [67] as well this methodology spans intraspecific and interspecific evolution and provides evidence fundamental to inferring the process of speciation [68], [69]. Others species delimitation methods could be analyzed, such as [17], [69]. However, we did diverse analyses and obtained consistent results.

Distance analyses showed that the intraspecific genetic variation within M. olfersii is lower than the interspecific variation [27]. This divergence is strongly based on three genes; however, the COI gene shows the most difference among the closest species, corroborating the utility of the COI gene as a good marker to separate close and sibling species. Moreover, within the M. olfersii clade illustrated in all phylogenetic analyses, there is no genetic structure and no haplotype fixed in a single population. Therefore, the observed morphological variability must be phenotypic plasticity [10]. In the haplotype network, this condition is supported by haplotype sharing (polymorphic), showing a continuous gene flow among Caribbean and Brazilian populations.

M. olfersii×M. digueti

From the COI datasets, we found that M. olfersii remains in a single clade in ML and BI analyses (Fig. 3). However, analyses based on 16S datasets (Fig. 2) and on concatenated genes, the sequence of M. olfersii retrieved from GenBank (ID: AY377849) was not separated from M. digueti branch. Although this specimen had been not analyzed, these results suggest that the identification this exemplar is incorrect. On H3 datasets indicated that close congener species appeared inside the M. olfersii clade (Fig. 4). Besides, this intimate relationship would be related to the condition of sibling species between M. olfersii and M. digueti [9].

Macrobrachium olfersii occurs on Atlantic and M. digueti occurs on Pacific coast of Americas. They are closely related and there are a small number of differences morphological between them. Apart from the condition of sibling species, they are considered cryptic species (Rossi & Mantelatto, in preparation). Their geographic distribution, the obtained divergence by COI and 16S, the phylogenies based on COI analysis confirm the valid of both species. The results by H3 showed the high genetic similarities, due to be more conservative gene. Although we did not clock molecular analysis, we suggested that the split of these species could be associated with closure of the Isthmus of Panama due these genes have lower rates of evolution and the pattern of the geographic distribution.

It is possible that these taxa have radiated following the closure of the Isthmus of Panama, similarly to other decapods such as the snapping shrimps [70]. In contrast, some studies have indicated that some cryptic species-complexes of shrimps show older (pre-Isthmian) divergences that were probably responses to environmental changes [71], [72], [73]. Consequently, these closely related sibling species cannot be separated by using only conservative genes. The results by H3 nDNA were insufficient to show the difference between these cryptic species. Although using concatenated genes it was not visualized, we suggest the need for future phylogenetic studies using other lines of evidence (such as larval morphology) and larger numbers of specimens to improve our knowledge of the natural relationships between this pair of species.

M. olfersii×M. faustinum

Another fascinating result was the relative positions of M. olfersii and M. faustinum, which were located on separate branches on all phylogenetic analyses (Figs. 2, 3 and 5), but H3 (Fig. 4), which is too conservative gene. Distance analyses showed the percentage of intraspecific was lower than interspecific variation. M. olfersii data ranged from 0.00 to 0.18% for 16S, 0.00 to 0.95% for COI, and 0.00 to 0.27% for H3. The obtained values between M. olfersii and M. faustinum was 3.84% for 16S, ranging from 12.51 to 12.70% for COI and from 0.27 to 0.54% for H3. In previous study, Macrobrachium analyzed ranged 15.6% (M. americanum and M. nattereri) for 16S and 2.3 (M. americanum and M. carcinus) to 20.6% (M. heterochirus and M. borelli) for COI. In the same study it was showed an intraspecific divergence ranging from 0 to 0.9 for COI from different M. olfersii specimens [27].

This fact disagrees with previous morphological studies that suggested that M. faustinum is a junior synonym of M. olfersii, because of the close similarity between them [74], [39], [9], [11]. The voucher specimen that was deposited in GenBank as M. faustinum (ID: HM352461) was analyzed, and its identification is incorrect. Therefore, the sequence for “M. faustinum” available in GenBank should be redesignated as M. olfersii. This case proves the importance to verify the identification of the specimens before the molecular study and the necessity of the existence of voucher.

Our result is also supported by the geographical distributions of these species. M. faustinum occurs only in the West Indies and Florida [9], [75], [76], but the specimen available in GenBank was from Curarigua, Venezuela (IVIC 1083). M. olfersii does not occur in the West Indies [75].

Although the phylogenetic topology obtained here was well supported, another study has been conducted (Rossi & Mantelatto, in preparation) with larger numbers of specimens from some regions and including other related species, in order to complete the entire set of species and improve knowledge of the phylogenetic relationships of the M. olfersii complex.

Gene Flow and Dispersion

Freshwater prawns of the genus Macrobrachium are thought to have originated from marine ancestors in the beginning of the Pleistocene, some of which subsequently migrated towards freshwater in more than one time [77], [64]. A colonization of freshwater is considered as the invasion by the ancestor of a lineage followed by the subsequent diversification of that lineage within continental waters [78]. These species acquired many physiological, ecological and behavioral adaptations [77], [79], [6]. Three basic types of larval developmental patterns can be recognized in this genus there are several species showing transitional developmental models. Some of them need estuarine and freshwater to complete their life cycle, which implicates in recurring migrations between both environments, such as M. olfersii [80], [6].

Molecular variation among the groups and the populations was not significant, again confirming the occurrence of a continuous gene flow. Consequently, the possibility of differentiation at the genetic level is decreased [81]. The existence of gene flow is plausible, since the M. olfersii larvae could be carried out on aquatic plants or associated with cultured species, as may have been the means of introduction of M. olfersii into Florida, United States [13]. On the other hand, the species may also have used favorable currents to reach Florida. This latter argument is strengthened by the finding of M. olfersii populations in all east-coast drainages of Florida [76].

Following the idea of M. olfersii to be undergoing a continuous process of adapting completely for freshwater environments [77], [32], [64], because it still depends on brackish water, Larvae of Macrobrachium can survive in high concentrations (over 30 ppt) of salt water [82], [83], [80]. Macrobrachium olfersii has been found in natural estuarine habitats at salinities up to 36 ppt. This finding indicates that the species has a higher salinity tolerance than was suggested by earlier laboratory studies [83], and it can survive in sea water for extend periods of time [76].

At the same time, studies of the physiology of freshwater shrimps have showed that although the Na+/K+ -ATPase systems of the animals do not function at maximal activity, the Na+ transport systems respond to salt loading when they are in the process of acclimating to high salinity [79]. We known that larvae of Macrobrachium species have a widely dispersion [11], [13], [83]. Marine current could carry M. olfersii larvae over long distances, from southern Brazil to the Caribbean or the contrary, allowing a continuous gene flow.

Conclusions

In spite of its high morphological variability and wide geographic distribution, Macrobrachium olfersii has a lower intraspecific genetic divergence than interspecific. Phylogenetic analysis based on COI mtDNA sequences revealed surely that M. olfersii forms a monophyletic clade. Also, there were no differences among the populations. This result confirms that continuous gene flow exists among Caribbean and Brazilian populations of M. olfersii, as shown by the haplotype net, and supports the characterization of these populations as a single species. Macrobrachium digueti is a sister group, from the Pacific coast.

Moreover, the analysis of 16S mtDNA and H3 nDNA sequences provided evidence that M. olfersii has recently diverged from other Macrobrachium species from Central America, namely the M. olfersii complex. In conclusion, our findings confirm using H3 nDNA and 16S rDNA sequences as molecular markers for separated species identification whereas the COI gene is better suited to address genetic lineages and to explore possible cryptic species. However, addition of data from more specimens and other species is required to enhance the confirmation and resolution of the phylogenetic relationships of these groups. Morphological studies of the species belonging to this complex are in progress, to complete and elucidate this scenario.

Acknowledgments

We are deeply grateful to many colleagues and friends (Alexandre Almeida, Darryl Felder, Emerson Mossolin, Fernando Alvarez, Georgina B. Buckup, Ingo Werthmann, José Luis Villalobos, Juan Bolaños, Luis Ernesto Arruda Bezerra, Marcos Tavares, Michael Tuerkay, Peter Schwendinger, and Rodrigo Johnsson) for help in collecting and lending material from collections. Special thanks are due to Leonardo Pileggi, Mariana Terossi, Rafael Robles, other members of LBSC, and Alvaro Costa for help in collecting, making available some essential fresh specimens, and for critical discussions during the preparation of this manuscript. Thanks also to Dr. Janet Reid (JWR Associates) for providing the English review service and the anonymous reviewers for their suggestions and contributions towards improving this paper. The support and assistance of the Postgraduate Program in Comparative Biology of FFCLRP/USP are gratefully acknowledged.

Author Contributions

Conceived and designed the experiments: NR FLM. Performed the experiments: NR FLM. Analyzed the data: NR FLM. Contributed reagents/materials/analysis tools: NR FLM. Wrote the paper: NR FLM.

References

  1. 1. Short JW (2004) A revision of Australian river prawns, Macrobrachium (Crustacea: Decapoda: Palaemonidae). Hydrobiologia 525(1–3): 1–100 .
  2. 2. Munansinghe DHN (2010) Phylogenetic positions of some species of the genus Macrobrachium Bate, 1868 (Crustacea, Decapoda, Palaemonidae) in Sri Lanka. Journal of the National Science Foundation of Sri Lanka 38(3): 193–199.
  3. 3. Wowor D, Muthu V, Meier R, Balke M, Cai Y, et al. (2009) Evolution of life history traits in Asian freshwater prawns of the genus Macrobrachium (Crustacea: Decapoda: Palaemonidae) based on multilocus molecular phylogenetic analysis. Molecular Phylogenetics and Evolution 52(1): 340–350 .
  4. 4. De Grave S, Fransen CHJM (2011) Carideorum catalogus: the Recent species of the dendrobranchiate, stenopodidean, procaridean and caridean shrimps (Crustacea: Decapoda). Zoologische Mededelingen 89(5): 195–589 Available: http://www.zoologischemededelingen.nl/85/nr02/a01.
  5. 5. 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(4): 622–632 .
  6. 6. Bauer RT (2011) Amphidromy and migrations of freshwater shrimps. I. Costs, benefits, evolutionary origins and an unusual case of amphidromy. In: Akira Asakura et al. (Eds.) New frontiers in crustacean biology. Koninklijke Brill NV, Leiden, 145–156.
  7. 7. Hedgpeth JW (1949) The North American species of Macrobrachium (river shrimp). Texas Journal of Science 1(3): 28–38.
  8. 8. Bowles DE, Aziz K, Knight CL (2000) Macrobrachium (Decapoda: Caridae: Palaemonidae) in the contiguous United States: A review of the species and assessment of threats to their survival. Journal of Crustacean Biology 20(1): 158–171.
  9. 9. 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.
  10. 10. Villalobos FA (1969) Problemas de especiación en América de un grupo de Palaemonidae del genero Macrobrachium. Proceedings of the World Scientific Conference on the Biology and Culture of Shrimps and Prawns. FAO Fisheries Reports 57(3): 1055–1066.
  11. 11. Holthuis LB, Provenzano AJ (1970) New distribution records for species of Macrobrachium with notes on the distribution of the genus in Florida (Decapoda, Palaemonidae). Crustaceana 19(2): 211–213.
  12. 12. Ortmann AE (1897) Os camarões de água doce da América do Sul. Revista do Museu Paulista 2: 173–216.
  13. 13. Anderson G, Filingame JA (1980) The occurrence of Macrobrachium olfersii (Wiegmann, 1836) and Macrobrachium carcinus (Linnaeus, 1758) in Southern Mississippi, U.S.A. (Decapoda, Palaemonidae). Crustaceana 39(1): 90–94.
  14. 14. Williams AB (1984). Shrimps, lobsters and crabs of the Atlantic coast of the eastern United States, Maine to Florida, Smithsonian Institution Press, Washington, 550pp.
  15. 15. Melo GAS (2003) Famílias Atyidae, Palaemonidae e Sergestidae. In: Melo, GAS Manual de identificação dos Crustacea Decapoda de água doce do Brasil. Editora Loyola, p. 289–415.
  16. 16. Schwander T, Leimar O (2011) Genes as leaders and followers in evolution. Trends in Ecology and Evolution 26(3): 143–151 .
  17. 17. Yang Z, Ranala B (2010) Bayesian species delimitation using multilocus sequence data. Proceedings of the National Academy of Science of the United States of America 107(20): 9264–9269 Available: www.pnas.org/cgi/doi/10.1073/pnas.0913022107.
  18. 18. Vergamini FG, Pileggi LG, Mantelatto FL (2011) Phylogenetic analysis and genetic variability of the Amazon river prawn Macrobrachium amazonicum (Decapoda, Caridea, Palaemonidae) from inland and coastal waters. Contributions to Zoology 80(1): 67–83.
  19. 19. Barros MP (1995) Dados biológicos sobre Macrobrachium olfersii (Wiegmann, 1836) (Decapoda: Palaemonidae) da praia da Vigia, Garopava, Santa Catarina, Brasil. Biociências 3(2): 239–252.
  20. 20. Mossolin EM, Bueno SLS (2002) Reproductive biology of Macrobrachium olfersi (Decapoda, Palaemonidae) in São Sebastião, Brazil. Journal of Crustacean Biology 22(2): 367–376 .
  21. 21. Mossolin EM, Bueno SLS (2003) Relative growth of the second pereiopod in Macrobrachium olfersi (Wiegmann, 1836) (Decapoda, Palaemonidae). Crustaceana 76(3): 363–376 .
  22. 22. Müller YMR, Nazari EM, Simões-Costa MS (2003) Embryonic stages of the freshwater prawn Macrobrachium olfersii (Decapoda, Palaemonidae). Journal of Crustacean Biology 23(4): 869–875 Available: http://dx.doi.org/10.1651/C-2408.
  23. 23. Kuris AM, Ra’Anan Z, Sagi A, Cohen D (1987) Morphotypic differentiation of male Malaysian giant prawns, Macrobrachium rosenbergii. Journal of Crustacean Biology 7(2): 219–237 .
  24. 24. Moraes-Riodades PMC, Valenti WC (2004) Morphotypes in male Amazon river prawns, Macrobrachium amazonicum. Aquaculture 236(1–4): 297–307 .
  25. 25. Wortham JL, Maurik LN (2012) Morphology and morphotypes of the hawaiian river shrimp, Macrobrachium grandimanus. Journal of Crustacean Biology 32(4): 545–556 .
  26. 26. Ammar D, Müller YMR, Nazari EM (2001) Biologia reprodutiva de Macrobrachium olfersii (Wiegmann) (Crustacea: Decapoda: Palaemonidae) coletados na ilha de Santa Catarina, Brasil. Revista Brasileira de Biologia 18(2): 529–537.
  27. 27. 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(1): 194–208
  28. 28. Pileggi LG, Mantelatto FL (2012) Taxonomic revision of doubtful Brazilian freshwater shrimp species of genus Macrobrachium (Decapoda, Palaemonidae). Iheringia, Série Zoologia, 102(4): In press.
  29. 29. Murphy NP, Austin CM (2004) Multiple origins of the endemic Australian Macrobrachium (Decapoda: Palaemonidae) based on 16S rRNA mitochondrial sequences. Australian Journal of Zoology 52(5): 549–559 .
  30. 30. Rintelen KV, Page TJ, Cai Y, Roe K, Stelbrink B, et al. (2012) Drawn to the dark side: A molecular phylogeny of freshwater shrimps (Crustacea: Decapoda: Caridea: Atyidae) reveals frequent cave invasions and challenges current taxonomic hypotheses. Molecular Phylogenetics and Evolution 63: 82–96 .
  31. 31. 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 .
  32. 32. Liu M, Cai Y, Tzeng C (2007) Molecular systematics of the freshwater prawn genus Macrobrachium, Bate, 1868 (Crustacea: Decapoda: Palaemonidae) inferred from mtDNA sequences, with emphasis on East Asian species. Zoological Studies 46(3): 272–289.
  33. 33. Baker N, Bruyn M, Mather PB (2008) Patterns of molecular diversity in wild stocks of the redclaw crayfish (Cherax quadricarinatus) from Northern Australia and Papua New Guinea: impacts of Plio-Pleistocene landscape evolution. Freshwater Biology 53(8): 1592–1605 .
  34. 34. Schubart CD (2009) Mitochondrial DNA and Decapod Phylogenies: The Importance of Pseudogenes and Primer Optimization In: Martin JW, Crandall KA, Felder DL (Eds), Crustacean Issues: Decapod Crustacean Phylogenetics. CRC Press, Boca Raton, 41–65.
  35. 35. 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 .
  36. 36. Murphy NP, Austin CM (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): 184–197 .
  37. 37. Bracken HD, Toon A, Felder DL, Martin JW, Finley M, et al. (2009) The decapod tree of life: compiling the data and moving toward a consensus of decapods evolution. Arthropod Systematics & Phylogeny 67(1): 99–116.
  38. 38. Valencia DM, Campos MR (2007) Freshwater prawns of the genus Macrobrachium Bate, 1868 (Crustacea: Decapoda: Palaemonidae) of Colombia. Zootaxa 1456: 1–44.
  39. 39. Sawaya MP (1946) Sôbre alguns camarões d’água doce do Brasil. Boletim da Faculdade de Filosofia, Ciências e Letras da Universidade de São Paulo, Zoologia 11: 393–407.
  40. 40. Mantelatto FL, Robles R, Felder DL (2007) Molecular phylogeny of the western Atlantic species of the genus Portunus (Crustacea: Brachyura, Portunidae). Zoological Journal of the Linnean Society 150(1): 211–220 .
  41. 41. Mantelatto FL, Pardo LM, Pileggi LG, Felder DL (2009a) Taxonomic re-examination of the hermit crab species Pagurus forceps and Pagurus comptus (Decapoda, Paguridae) by molecular analysis. Zootaxa 2133: 20–32.
  42. 42. Mantelatto FL, Robles R, Schubart CD, Felder DL (2009b) Molecular phylogeny of the genus Cronius Stimpson 1860, with reassignment of C. tumidulus and several American species of Portunus to the genus Achelous de Haan, 1833 (Brachyura: Portunidae). In: Martin JW, Crandall KA, Felder DL (Eds), Crustacean Issues: Decapod Crustacean Phylogenetics. CRC Press, Boca Raton, 567–579.
  43. 43. Porter ML, Pérez-Losada M, Crandall KA (2005) Model-based multi-locus estimation of decapod phylogeny and divergence times. Molecular Phylogenetics and Evolution 37(2): 355–369 .
  44. 44. Page TJ, Cook BD, Rintelen T, Rintelen K, Hughes JM (2008) Evolutionary relationships of atyid shrimps imply both ancient Caribbean radiations and common marine dispersals. Journal of the North American Benthological Society 27(1): 68–83
  45. 45. Colgar DJ, McLauchlan A, Wilson GDF, Livingston SP, Edgecombe GD, et al. (1998) Histone H3 and U2 snRNA DNA sequences and arthropod molecular evolution. Australian Journal of Zoology 46: 419–437
  46. 46. Crandall KA, Fitzpatrick Jr JF (1996) Crayfish molecular systematics: using a combination of procedures to estimate phylogeny. Systematic Biology 45: 1–26 .
  47. 47. Palumbi SR, Benzie J (1991) Large mitochondrial DNA differences between morphologically similar penaeid shrimp. Molecular Marine Biology and Biotechnology 1(1): 27–34.
  48. 48. Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41: 95–98.
  49. 49. 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 .
  50. 50. Kumar S, Dudley J, Nei M, Tamura K (2008) MEGA: A biologist-centric software for evolutionary analysis of DNA and protein sequences. Briefings in Bioinformatics 9: 299–306.
  51. 51. Song H, Buhay JE, Whiting MF, Crandall KA (2008) Many species in one: DNA barcoding overestimates the number of species when nuclear mitochondrial pseudogenes are coamplified. Proceedings of the National Academy of Science of the United States of America 105(36): 13486–13491 www.pnas.org/cgi/doi/10.1073/pnas.0803076105.
  52. 52. Swofford DL (2003) PAUP: Phylogenetic analysis using parsimony (and other metods). Version 4.0b.10. Sunderland: Sinauer Associates.
  53. 53. Posada D, Buckley T (2004) Model selection and model averaging in phylogenetics: advantages of Aikaike information criterion and Bayesian approaches over likelihood ratio tests. Systematic Biology 53: 793–808 .
  54. 54. Posada D, Crandall KA (1998) ModelTest: testing the model of DNA substitution. Bioinformatics 14(9): 817–818 .
  55. 55. Felsenstein J (1985) Confidence limits on phylogenies: an approach using the bootstrap. Evolution 39(4): 783–791 .
  56. 56. Huelsenbeck JP, Ronquist F (2001) MrBayes: Bayesian inference of phylogeny. Bioinformatics 17: 754–755.
  57. 57. Ranbaut A (2007) FigTree. http://tree.bio.ed.ac.uk/software/figtree/.
  58. 58. Rozas J Rozas R (1999) DnaSP version 3.0: an integrated program for molecular population genetics and molecular evolution analysis. Bioinformatics 15: 174–175 .
  59. 59. Excoffier L, Laval G, Schneider S (2005) Arlequin (version 3.0): an integrated software package for population genetics data analysis. Evolutionary Bioinformatics Online 1: 47–50.
  60. 60. Clement M, Posada D, Crandall KA (2000) TCS: a computer program to estimate gene genealogies. Molecular Ecology 9: 1657–1660 .
  61. 61. Bandelt HJ, Forster P, Röhl A (1999) Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution 16: 37–48.
  62. 62. Excoffier L, Smouse PE, Quattro JM (1992) Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics 131: 479–491.
  63. 63. Ng PKL, Shih HT, Naruse T, Shy JY (2010) Using Molecular Tools to Establish the Type Locality and Distribution of the Endemic Taiwanese Freshwater Crab Geothelphusa chiui Minei, 1974 (Crustacea: Brachyura: Potamidae), with Notes on the Genetic Diversity of Geothelphusa from Eastern Taiwan. Zoological Studies 49(4): 544–555.
  64. 64. Ashelby CW, Page TJ, De Grave S, Hughes JM, Johnson ML (2012) Regional scale speciation reveals multiple invasions of freshwater in Palaemoninae (decapod). Zoologica Scripta 41(3): 293–306 .
  65. 65. Robles R, Tudge C C, Dworschak PC, Poore GCB, Felder DL (2009) Molecular Phylogeny of the Thalassinidea Based on Nuclear and Mitochondrial Genes In: Martin JW, Crandall KA, Felder DL (Eds), Crustacean Issues: Decapod Crustacean Phylogenetics. CRC Press, Boca Raton, 309–326.
  66. 66. Schnabel KE, Ahyong ST, Maas EW (2011) Galatheoidea are not monophyletic – Molecular and morphological phylogeny of the squat lobsters (Decapoda: Anomura ) with recognition of a new superfamily. Molecular Phylogenetics and Evolution 58: 157–168 .
  67. 67. Mishler BD, Theriot E (2000) The phylogenetic species concept sensu Mishler and Theriot: monophyly, apomorphy, and phylogenetic species concepts. In: Wheeler QD, Meier R (Eds.), Species concepts and phylogenetic theory: a debate. Columbia University Press, NY. [Three chapters] Pp.44–54, 119–132, 179–184.
  68. 68. Templeton AR (1994) The role of molecular genetics in speciation studies. In: Schierwater B, Streit B, Wagner GP, DeSalle R (Eds) Molecular Ecology and Evolution: Approaches and Applications, Birkha-user-Verlag, Basel, 455–477.
  69. 69. Ence DD, Carstens BC (2011) SpedeSTEM: a rapid and accuratemethod for species delimitation. Molecular Ecology Resources 11: 473–480 .
  70. 70. Mathews LM (2006) Cryptic biodiversity and phylogeographical patterns in a snapping shrimp species complex. Molecular Ecology 15: 4049–4063 .
  71. 71. 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.
  72. 72. Knowlton N, Weigt LA (1998) New dates and new rates for divergence across the Isthmus of Panama. Proceedings of the Royal Society, Biological Sciences 265: 2257–2263 .
  73. 73. 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 .
  74. 74. Rathbun MJ (1900) Results of the Branner-Agassiz Expedition to Brazil I. The decapod and stomatopod Crustacea. Proceedings of the Washington Academy of Sciences 2: 133–156.
  75. 75. Chace FA, Hobbs Jr HH (1969) The freshwater and terrestrial decapod crustaceans of the West Indies with special reference to Dominica. Bulletin of the United States National Museum 292: 1–258.
  76. 76. 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 .
  77. 77. Jalihal DR, Sankolli KN, Shenoy S (1993) Evolution of larval development and the process of freshwaterization in the prawn genus Macrobrachium Bate, 1868 (Decapoda, Palaemonidae). Crustaceana 65: 365–376.
  78. 78. Boxshall GA, Jaume D (2000) Making waves: the repeated colonization fresh water by copepod crustaceans. Advances in Ecological Research 31: 61–79.
  79. 79. McNamara JC, Torres AH (1999) AID-JEZ3>3.0.CO;2-V.
  80. 80. Gamba AL, Rodríguez G (1987) Migratory behavior of postlarval white, Penaeus schmitti, and river shrimps, Macrobrachium olfersi and Macrobrachium acanthurus in their zone of overlap in a tropical lagoon. Bulletin of Marine Science 40(3): 454–463.
  81. 81. Slatkin M (1987) Gene flow and the geographic structure of natural populations. Science 236: 787–792 .
  82. 82. Choudhury PC (1970) Complete larval development of the Palaemonidae shrimp Macrobrachium acanthurus (Wiegmann, 1836) reared in the laboratory. Crustaceana 18(2): 113–132.
  83. 83. Gamba AL (1982) Macrobrachium: its presence in estuaries of the Northern Venezuelan coast (Decapoda, Palaemonidae). Caribbean Journal of Science 18: 1–4.