Skip to main content

Genomic organization of repetitive DNAs highlights chromosomal evolution in the genus Clarias (Clariidae, Siluriformes)

Abstract

Background

The genus Clarias (Clariidae, Siluriformes) contains at least 61 species naturally spread over vast regions of Asia, India and Africa. However, Clarias species have also been introduced in many different countries and represent the most widespread catfishes in the world. These fishes are also known as “walking catfishes” due to their ability to move over land. A large degree of chromosomal variation has been previously found in this family, mainly using conventional cytogenetic investigations, with diploid chromosome numbers ranging between 48 and 100. In this study, we analyzed the karyotype structure and distribution of four repetitive DNA sequences (5S and 18S rDNAs and (CA)15 and (GA)15 microsatellites) in three Clarias species (C. batrachus, C. gariepinus, C. macrocephalus), as well as in a probable natural hybrid of the two latter species from different Thailand river basins.

Results

Clarias gariepinus and C. macrocephalus had 2n = 56 and 2n = 54, respectively, as well as karyotypes composed mainly by metacentric and submetacentric chromosomes. Their karyotypes differed in the number and location of 5S and 18S rDNA sites and in the degree of microsatellite accumulation. An intermediate chromosomal pattern incorporating those of the parental species was found in the probable hybrid, confirming its interspecific origin. Clarias batrachus had 2n = 104 chromosomes and its karyotype was dominated by mainly acrocentric elements, indicating that unusual multiple centric fissions were involved in its karyotype differentiation. The karyotype of this species presented an unexpected dispersion of ribosomal DNAs, possessing 54 and 12 sites of 5S and 18S rDNAs, respectively, as well as a high accumulation and differential distribution of both microsatellite repeats, representing ‘hot spots’ for chromosomal rearrangement.

Conclusion

Both conventional and molecular cytogenetic markers were useful tools for demonstrating remarkable evolutionary dynamism and highlighting multiple chromosomal rearrangements and hybridization events correlated with the notable karyotypic diversity of these walking catfishes.

Background

The catfish family Clariidae comprises 14 genera and more than 115 species found in India, Syria, Southern Turkey, Southeast Asia and Africa, with the highest species diversity found in the latter [1]. The genus Clarias Scopoli, 1763, comprises at least 57 species and is widely distributed across Africa and Southeast Asia [2, 3]. These fishes are known as “walking catfishes” because they have an accessory air-breathing organ that allow them to survive for months in oxygen-poor water or even completely out of water [4]. Moreover, some species are able to migrate over land using wriggling movements. These characteristics and behaviour may affect the dispersal, speciation and genetic parameters of the population, potentially altering the processes from those commonly recognized in other obligatory freshwater fishes. In Clarias species examined to date, diploid chromosome numbers range between 2n = 48 and 2n = 56, with the exception of C. pachynema (2n = 66) and one population of C. batrachus (2n = 100) (Table 1). However, all chromosomal data were obtained by conventional Giemsa-stained chromosomes and molecular cytogenetic studies are still virtually absent.

Table 1 Review of available data on 2n, karyotypes and sex systems in the genus Clarias

The molecular organization and cytogenetic mapping of repetitive DNA elements, including satellites, multigene families and microsatellite repeats, have been analyzed in a large number of species and have demonstrated enormous potential for expanding our understanding of karyotype differentiation in fishes (reviewed in [5]). In fact, the correlation between the presence of high amounts of repetitive sequences and the higher number of chromosomal rearrangements has been widely demonstrated, since the accumulation of repetitive DNAs in particular genomic regions may prompt chromosome breakages, deletions, inversions and amplifications [6].

In this study, we analyzed the karyotype structure and distribution of four repetitive DNA sequences [5S and 18S rDNAs and (CA)15 and (GA)15 microsatellites] in three Clarias species (C. gariepinus, C. macrocephalus and C. batrachus) and in a probable natural hybrid of C. gariepinus and C. macrocephalus from different Thailand river basins with the aim of investigating their chromosomal differentiation and relationships. We observed remarkable chromosomal dynamism and karyotype characteristics that confirmed the hybridization of C. gariepinus x C. macrocephalus.

Results

Karyotypes

No differences between male and female karyotypes were observed in any species. Clarias gariepinus showed a chromosome number of 2n = 56 (18 m + 20 sm + 18 st/a) and a fundamental number (NF) equal to 94. Clarias macrocephalus showed 2n = 54 (18 m + 20 sm + 16 st/a) and a NF of 92. The probable natural hybrid of C. gariepinus x C. microcephalus had a 2n between the two species (55) with a karyotype composed of 18 m + 20 sm + 17 st/a, and a NF equal to 91 (Fig. 1). Clarias batrachus showed a 2n = 104 (2 m + 4 sm + 98 st/a), with the NF equal to 110 (Fig. 2).

Fig. 1
figure 1

Karyotypes arranged from metaphase chromosomes of Clarias gariepinus (2n = 56), Clarias macrocephalus (2n = 54) and the natural hybrid of these species (2n = 55) after Giemsa staining and FISH with 18S rDNA (red) and 5S rDNA (green) probes. Note the intermediate level of distribution of the ribosomal sites in the hybrid specimen compared to the parental species. Scale bar = 5 μm

Fig. 2
figure 2

Karyotypes arranged from metaphase chromosomes of Clarias batrachus (2n = 104) after Giemsa staining and FISH with 18S rDNA (red), 5S rDNA (green), and (GA)15 and (CA)15 microsatellite probes. Note the high dispersion of 5S rDNA sites in the karyotype. Scale bar = 5 μm

Chromosome mapping of 5S and 18S rDNA sequences

The 18S rDNA probe hybridized to the subtelomeric/telomeric region of one medium-sized sm chromosomal pair in C. gariepinus and in one large st/a pair in C. macrocephalus. The 5S rDNA sequences were located in two small st/a pairs in C. gariepinus and in only one st/a pair in C. macrocephalus. Accordingly, the supposed hybrid of C. gariepinus x C. microcephalus had the intermediate number for both rDNA probes, with two 18S rDNA sites present in non-homologous chromosomes, in addition to three 5S rDNA sites (Fig. 1).

However, C. batrachus showed a surprising increase in the number of both rDNA classes, with six chromosomal pairs harbouring 18S rDNA sites and 27 chromosomal pairs harbouring 5S rDNA sites, including a syntenic condition in one pair (Fig. 2).

Chromosome mapping of microsatellite sequences

In C. gariepinus, faint hybridization signals were found for these sequences across all chromosomes. However, in C. macrocephalus and C. batrachus, both (CA)15 and (GA)15 microsatellites were highly accumulated along all chromosomes. Significantly, the likely interspecific hybrid had the exact intermediate pattern of C. gariepinus and C. macrocephalus, with 27 chromosomes presenting a strong hybridization pattern, and the other 28 presenting weak accumulation of these sequences (Figs. 2 and 3).

Fig. 3
figure 3

Karyotypes arranged from metaphase chromosomes of Clarias gariepinus (2n = 56), Clarias macrocephalus (2n = 54) and the natural hybrid of these species (2n = 55) after FISH with (GA)15 and (CA)15 microsatellite probes. Note the weak distribution of both microsatellites in C. gariepinus, their strong accumulation in C. macrocephalus and the intermediate distribution pattern in the hybrid specimen. Scale bar = 5 μm

Discussion

Karyotype variability among Clarias species

Clariidae is a very well defined monophyletic family based on the presence of a unique arborescent suprabranchial organ that enables the species to breathe atmospheric oxygen [4]. Nevertheless, Clariidae species are remarkable for the considerable variation in their external morphology [79]. Therefore, the chromosomal divergence among Clarias species (Table 1) also parallels the morphological differentiation of the clariid catfishes. In fact, the remarkable variability of the 2n and NF values in different Clarias species indicates that distinct chromosomal rearrangements occur during the evolution of their karyotypes. Karyotypes and other chromosomal characteristics of the three species in this study confirmed the patterns found for other Clarias species, except C. batrachus (2n = 104). However, Clarias and related cytogenetic parameters warrant deeper discussion.

The walking catfish C. batrachus is native to Southeast Asia, but has been introduced outside its native range, where it is considered an invasive species responsible for invading aquaculture farms and preying on fish stocks [10, 11]. Indeed, C. batrachus has type locality in Java [12], while the populations from Indochina either represent introduced stocks or belong to other Clarias species, as a number of different species have been recently identified both in Indochina and in the Sunda islands [13]. Accordingly, cytogenetic data also point to very distinct karyotypes for this species: i) 2n = 50 in Malaysia, ii) 2n = 50–54 in India, iii) 2n = 56 in China and 2n = 100–104 in Thailand, in addition to sex and B chromosomes in some of these populations (Table 1). This brief overview of species suggests that the C. batrachus in Thailand, with 2n = 104 and a karyotype dominated by one-armed chromosomes, may represent a different unnamed species.

The hypothetical 2n for Siluriformes, as described in studies of different species of this order, was proposed to be 2n = 56, with a karyotype composed mainly by m-sm chromosomes [1416]. Chromosomal studies of species in the group Heteropneustidae, which is phylogenetically considered a sister-group to Clariidae [17], report that most of its members also have 2n = 56 [18]. Clarias gariepinus, with 2n = 56 chromosomes, as well as a higher number of two-armed chromosomes and few acrocentric chromosomes, retains the karyotype considered basal for Siluriformes. These data support the phylogeny proposed for this family based on mtDNA analysis [9], in which C. gariepinus is placed together with C. anguillaris, as both species contain 2n = 56 chromosomes [19]. The decrease in the 2n of other Clariidae species (such as C. macrocephalus - Fig. 2), suggests that chromosomal fusions also participated in the karyotypic differentiation in the family.

Phylogenetically the species C. batrachus shows a derived position in the family [9]. In the present study, this species presented an unusual 2n = 104 and a karyotype dominated by acrocentric chromosomes. The occurrence of such a high 2n could be indicative of a polyploidization event. However, when considering the large number of acrocentric chromosomes and their relatively small size compared to the chromosomes found in the other Clarias species, the present data suggest that multiple centric fissions are, in fact, the most plausible explanation for karyotype diversification in this species. A similar process culminating in an increased 2n number has also been reported for species of the genus Potamorhina. One species - P. altamazonica - with a 2n = 102, diverged from the most frequent 2n (54) found in other congeneric species through a process of multiple centric fissions. In that case, meiotic analysis showed only bivalents at metaphase I and confirmed a large scale occurrence of extensive chromosomal fissions [20]. Additionally, extensive centric fission and heterochromatinization have been proposed in the karyotype diversification of the Alaska black fish (Dallia pectoralis) [21].

The chromosomal distribution of repetitive DNA elements revealed remarkable differences among the analyzed species. Both C. gariepinus and C. macrocephalus presented two 18S rDNA sites but were located in distinct chromosomal pairs, while four 5S rDNA sites were present in C. gariepinus and only two in C. macrocephalus. In addition, analysis of C. batrachus revealed six chromosomal pairs harbouring 18S rDNA sites and 27 chromosomal pairs harbouring 5S rDNA sites, including a synteny case. Though rDNAs are among the most conservative components of the eukaryotic genome, undergoing minimal changes over hundreds of millions of years, this conservatism appears to be a powerful source for genome instability [22]. Due to high similarity among clusters, chromosomes that carry extended rDNA arrays could be involved in heterologous synapses and recombination [23], providing variations of these sites inside the karyotypes.

Hypervariability in the number and location of rDNA loci, as presently reported in C. batrachus, has been previously described for several groups [2426]. Variability in the number and position of rDNA sites suggested that chromosomal rearrangements played a role in the speciation of the plant Sideritis dendrochahorra [6]. This species possesses a large number of acrocentric chromosomes and multiple terminal 45S rDNA sites in most of its chromosomes. It has been suggested that in some groups, structural changes may be induced by selective pressures from ecological or environmental stresses [6, 27]. Among fishes, the spreading of rDNA has reportedly affected the recombination rates of two coexisting salmonid species, Coregonus albula and C. fontanae, leading to rapid genomic divergence and faster ecological speciation [28]. In some cases, transposable elements have been reported to play an important role in spreading rDNA sequences over the genome [24, 25]. Some classes of transposons appear to be able to “capture” entire genes and move them to other parts of the genome [29, 30]. Alternatively, several satellite DNA repeats may have originated from rDNA sequences and thus facilitate their dispersal into different genomic regions. For example in the fish Hoplias malabaricus, a highly amplified satellite repeat (5SHindIII-DNA) with sequence sharing similarity with 5S rDNA have been reported to exist in the centromeric region of several chromosomes [31, 32]. However, the reasons for the higher number of rDNA sites in C. batrachus still need to be clarified.

Repetitive DNAs as a powerful tool for Clarias hybrids identification

Cultured catfishes are one of the most important commodities in Thailand’s domestic freshwater fish market, where C. macrocephalus is always preferred for consumption due to its better taste [33]. However, this species has a very slow growth rate and low disease resistance. At the end of 1980s, hybrid catfish production increased, with the most produced fish derived by crossing the Asian catfish (C. macrocephalus) and the African catfish (C. gariepinus) [34]; fast growth and high disease resistance made these species attractive to farmers [33, 35]. However, this hybrid form is currently abundant in all Thailand’s rivers, threatening wild catfish populations due to competition, predation, and genetic introgression [34, 36]. In fact, these hybrids are potentially able to interbreed with the parental species, which can lead to gene pool introgression, as has been reported for C. macrocephalus [33].

Karyotype similarity, such as that which exists between C. gariepinus and C. macrocephalus, enhances the success of hybridization and back cross of many species [37]. In nature, the occurrence of chromosome numbers around the modal values of the clariid species may suggest that speciation within this group is related to a high rate of hybridization that results from common spawning [19].

However, when hybrids and individuals from parental species have similar karyotype structures, the use of differential cytogenetic techniques is required to provide distinguishable chromosomal markers [38]. Indeed, several known hybrids can be precisely identified and clearly distinguished from their parental species using cytogenetic markers [3840]. For example, conventional staining helped in the precise identification of the parental chromosomal types of the artificial hybrid resulted from the cross between Colossoma macropomum and Piaractus brachypomus [41]. In the present work, repetitive DNAs proved to be informative and allowed for precise characterization of the hybridization process. The hybrid C. gariepinus x C. macrocephalus had two 18S rDNA sites in non-homologous chromosomes in addition to three 5S rDNA sites; these were the exact intermediate numbers present in the parental species. In addition to the rDNA markers, microsatellites were highly useful for confirming the hybrid nature of the fish. These ubiquitous repeated sequences are present in all eukaryotic genomes, either in euchromatin or in heterochromatin, inside coding regions of structural genes or between other repetitive sequences [42]. In the current analyzed species, C. gariepinus presented faint hybridization signals of (CA)15 and (GA)15 at subtelomeric regions, while in C. batrachus and C. macrocephalus these sequences were highly accumulated in all chromosomes. Notably, the C. gariepinus x C. macrocephalus hybrid showed 27 chromosomes with the strong hybridization pattern, characteristic of C. macrocephalus chromosomes and the other 28 elements presenting a weak accumulation, found in C. gariepinus.

However, what are the genomic and ecological consequences of such hybrid unbalance? Interspecific hybrids not only led to diversification and speciation but also have important ecological consequences [43]. Some hybrids appear prevalent in nature, suggesting an evolutionary advantage for having different sets of chromosomes for adaptation and development [44]. It is known, for example, that hybridization can promote the activation of mobile elements and rapid genomic changes [45, 46]. Among fishes, interspecific hybrids between the red crucian carp (Carassius auratus) × common carp (Cyprinus carpio) showed faster genomic changes compared to the parental species, facilitated by intron gains and losses, homologous recombination and the formation of novel genes [47]. This ‘genomic shock’ has also been reported in many allopolyploid plants, translating as gene loss, chromosome mispairing, retrotransposon activation, altered methylation or rearrangements between parental genomes that could lead to novel gene sequences or differential homologous gene expression in hybrids throughout evolution [48].

Conclusions

This study demonstrated that both conventional and molecular cytogenetics were useful tools in highlighting the remarkable chromosomal diversification that characterizes evolution in the genus Clarias. The distribution of repetitive DNA sequences on chromosomes identified: i) a high variability in the number and position of rDNA sites, ii) multiple chromosomal rearrangements, including an unusual number of centric fissions, iii) a high dispersion of ribosomal sites in C. batrachus and iv) the natural hybridization of C. gariepinus x C. macrocephalus. In the latter case, the hybrid genome was fully identified by the distinctive patterns of microsatellites sequences found in both parental species. Concerning conservation issues, these data stressed that successful natural hybridizations are not limited by low variations in diploid number or conspicuous divergences in microsatellites distribution among species.

Methods

Materials

Individuals of both sexes of three Clarias species collected from different river basins of Thailand were analyzed: C. batrachus (8 ♂ and 8 ♀), C. gariepinus (8 ♂ and 8♀) and C. macrocephalus (9 ♂ and 8 ♀); also collected were those from a probable natural hybrid between C. gariepinus x C. microcephalus (7 ♂ and 9 ♀) (Fig. 4). The specimens were caught using a hand-net, placed in sealed plastic bags containing oxygen and clean water, and transported to the laboratory. Experiments were performed in accordance with ethical protocols, and anaesthesia using clove oil was administered prior to sacrificing the animals to minimize suffering. The process was approved by the Ethics Committee of Khon Kaen University and by the RGJ committee under no. PHD/K0081/2556. Mitotic chromosomes were obtained from cell suspensions of the anterior kidney using the conventional air-drying method [49]. The specimens were deposited in the fish collection of the Cytogenetic Laboratory, Department of Biology Faculty of Science, Khon Kaen University.

Fig. 4
figure 4

Collection sites of Clarias species from Thailand analyzed in the present study

Chromosome probes and FISH experiments

Two tandem-arrayed DNA sequences isolated from the genome of an Erythrinidae species, Hoplias malabaricus, were used as probes. The first probe contained a 5S rDNA repeat copy and included 120 base pairs (bp) of the 5S rRNA transcribed gene and 200 bp of the non-transcribed spacer (NTS) sequence [31]. The second probe corresponded to a 1,400 bp segment of the 18S rRNA gene obtained from nuclear DNA using PCR [32]. The 5S and 18S rDNA probes were cloned into plasmid vectors and propagated in DH5α Escherichia coli competent cells (Invitrogen, San Diego, CA, USA).

The 18S and 5S rDNA probes were labelled with Spectrum Orange-dUTP and Spectrum Green-dUTP, respectively, using nick translation according to the manufacturer’s recommendations (Roche, Mannheim, Germany).

The microsatellites (CA)15 and (GA)15, were used as probes and were synthesized as described by Kubat et al. [50]. These sequences were directly labelled with Cy3 at the 5’ terminus during synthesis by Sigma (St. Louis, MO, USA).

Fluorescence in situ hybridization (FISH) was performed in highly stringent conditions on mitotic chromosome spreads [51]. Metaphase chromosome slides were incubated with RNAse (40 μg/ml) for 1.5 h at 37 °C. After denaturation of the chromosomal DNA in 70 % formamide/2× SSC, pH 7.0, at 70 °C for 4 min, the hybridization mixture (2.5 ng/μl probes, 2 μg/μl salmon sperm DNA, 50 % deionized formamide, 10 % dextran sulphate) was dropped on the slides. The hybridization was performed overnight at 37 °C in a moist chamber containing 2× SSC. The first post-hybridization wash was performed with 2× SSC for 5 min at 65 °C, and a final wash was performed at room temperature in 1× SSC for 5 min. Finally, the slides were counterstained with DAPI and mounted in an antifade solution (Vectashield from Vector Laboratories).

Image processing

At least 30 metaphase spreads were analyzed to confirm the diploid chromosome number, karyotype structure and FISH results. Images were captured using an Olympus BX50 microscope (Olympus Corporation, Ishikawa, Japan) with CoolSNAP and Image Pro Plus 4.1 software (Media Cybernetics, Silver Spring, MD, USA). Chromosomes were classified as metacentric (m), submetacentric (sm), subtelocentric (st) or acrocentric (a) according to Levan et al. [52].

Abbreviations

2n:

diploid number

a:

acrocentric chromosome

DAPI:

4′,6-diamidino-2-phenylindole

dUTP:

2′-Deoxyuridine-5′-Triphosphate

FISH:

fluorescence in situ hybridization

FN:

fundamental number

m:

metacentric chromosome

NTS:

Non-transcribed spacer

PCR:

polymerase chain reaction

rDNA:

ribosomal DNA

rRNA:

ribosomal RNA

sm:

submetacentric chromosome

st:

subtelocentric chromosome

References

  1. Nelson JS. Fishes of the world. 4th ed. New York: John Wiley and Sons, Inc. press; 2006.

    Google Scholar 

  2. Skelton PH. A complete guide to the freshwater fishes of southern Africa. 2nd ed. Cape Town, South Africa: Struik Publishers; 2001.

    Google Scholar 

  3. Ng HH, Kottelat M. Clarias serniosus, a new walking catfish (Teleostei: Clariidae) from Laos. Zootaxa. 2014;3884(5):437–44.

    Article  PubMed  Google Scholar 

  4. Teugels GG, Adriaens D. Taxonomy and phylogeny of Clariidae: an overview. In: Arratia G, Kapoor BG, Chardon M, Diogo R, editors. Catfishes. NH: Science Publishers Inc, Enfield; 2003. p. 465–87.

    Google Scholar 

  5. Cioffi MB, Bertollo LAC. Chromosomal distribution and evolution of repetitive DNAs in fish. In: Garrido R, editor. Repetitive DNAs genome dynamics. Basel: Karger; 2012. p. 197–221.

    Google Scholar 

  6. Raskina O, Barber JC, Nevo E, Belyayev A. Repetitive DNA and chromosomal rearrangements: speciation related events in plant genomes. Cytogenet Genome Res. 2008;120:351–57.

    Article  CAS  PubMed  Google Scholar 

  7. Boulenger GA. A revision of the African silurid fishes of the subfamily Clariinae. Proceedings of the Zoological Society of London. 1907;1908b:1062–97.

    Google Scholar 

  8. Eyo JE. Congeneric discrimination of morphometric characters among members of the Pisces Genus Clarias (Clariidae) in Anambra River, Nigeria. The Zoologist. 2013;1:1–17.

    Google Scholar 

  9. Agnese JF, Teugels GG. Insight into the phylogeny of African Clariidae (Teleostei: Siluriformes): implications for their body shape evolution, biogeography and taxonomy. Mol Phylogenet Evol. 2005;36:546–53.

    Article  PubMed  Google Scholar 

  10. Teugels GG. The nomenclature of African Clarias species used in aquaculture. Aquaculture. 1984;38(4):373–4.

    Article  Google Scholar 

  11. Teugels GG. A systematic revision of the African species of the genus Clarias (Pisces; Clariidae). Ann Mus Roy Afr Centr Sci Zool. 1986;247:1–199.

    Google Scholar 

  12. Kottelat M. The fishes of the inland waters of Southeast Asia: a catalogue and core bibliography of the fishes known to occur in freshwaters, mangroves and estuaries. Raffles Bull Zool. 2013;27:1–663.

    Google Scholar 

  13. Ng HH, Kottelat M. The identity of Clarias batrachus (Linnaeus, 1758), with the designation of a neotype (Teleostei: Clariidae). Proceedings of the Linnaean Society of London. 2008;153:725–32.

    Article  Google Scholar 

  14. Ráb P. Karyotype of European catfish Silurus glanis (Siluridae, Pisces), with remarks on cytogenetics of siluroid fishes. Folia Zool Brno. 1981;32(2):271–86.

    Google Scholar 

  15. LeGrande WH. Chromosomal evolution in North American catfishes (Siluriformes, Ictaluridae) with particular emphasis on the madtoms, Noturus. Copeia. 1981;1:33–52.

    Article  Google Scholar 

  16. Oliveira C, Gosztonyi AE. A cytogenetic study of Diplomystes mesembrinus (Teleostei, Siluriformes, Diplomystidae) with a discussion of chromosome evolution in Siluriforms. Caryologia. 2000;53:31–7.

    Article  Google Scholar 

  17. Diogo R. Higher-level phylogeny of Siluriformes: an overview. In: Arratia G, Kapoor BG, Chardon M, Diogo R, editors. Catfishes Science. Publishers: Inc Enfield; 2003. p. 353–84.

    Google Scholar 

  18. Malla TM, Ganesh N. Cytogenetic and tissue toxicity by synthetic sindoor in freshwater catfish. Biomed Pharmacol J. 2009;2:885–9.

    Google Scholar 

  19. Eyo JE. Cytogenetic variations in Clarias species (Clariidae: Siluriformes) of the Anambra river using leucocytes culture tecniques. Anim Res Int. 2005;2(1):275–86.

    Google Scholar 

  20. Feldberg E, Porto JIR, Bertollo LAC. Karyotype evolution in Curimatidae (Teleostei, Characiformes) of the Amazon region. 1. Studies on the genera Curimata, Psectrogaster, Steindachnerina and Curimatella. Rev Brasil Genet. 1992;15(2):369–83.

    Google Scholar 

  21. Crossman EJ, Ráb P. Chromosome-banding study of the Alaska blackfish, Dallia pectoralis (Euteleostei: Esocae) with implications for karyotype evolution and relationships of esocoid fishes. Can J Zool. 1996;74:147–56.

    Article  Google Scholar 

  22. Rebordinos L, Cross I, Merlo A. High Evolutionary Dynamism in 5S rDNA of Fish: State of the Art. Cytogenet Genome Res. 2013;141:103–13.

    Article  CAS  PubMed  Google Scholar 

  23. Raskina O, Belyayev A, Nevo E. Quantum speciation in Aegilops: molecular cytogenetic evidence from rDNA cluster variability in natural populations. Proc Natl Acad Sci USA. 2004;101:14818–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Cioffi MB, Martins C, Bertollo LAC. Chromosome spreading of associated transposable elements and ribosomal DNA in the fish Erythrinus erythrinus. Implications for genome change and karyoevolution in fish. BMC Evol Biol. 2010;10:271.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Symonová R, Majtánová Z, Sember A, Staaks GBO, Bohlen J, Freyhof J, et al. Genome differentiation in a species pair of coregonine fishes: an extremely rapid speciation driven by stress - activated retrotransposons mediating extensive ribosomal DNA multiplications. BMC Evol Biol. 2013;13:42.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Cioffi MB, Bertollo LAC, Villa MA, Oliveira EA, Tanomtong A, Yano CF, et al. Genomic organization of repetitive DNA elements and its implications for the chromosomal evolution of channid fishes (Actinopterygii, Perciformes). Plos One. 2015;10(6):e0130199.

    Article  PubMed Central  Google Scholar 

  27. Cox AV, Abdelnour GJ, Bennett MD, Leitch IJ. Genome size and karyotype evolution in the Slipper orchids (Cypripedioideae: Orchidaceae). Am J Bot. 1998;85:681–7.

    Article  CAS  PubMed  Google Scholar 

  28. Pereira CS, Pazian MF, Ráb P, Collares-Pereira MJ. Dynamics of Rex3 in the genomes of endangered Iberian Leuciscinae (Cyprinidae) and their natural hybrids. Mol Cytogenet. 2015;8:81.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Jiang N, Bao Z, Zhang X, Eddy SR, Wessler SR. Pack-MULE transposable elements mediate gene evolution in plants. Nature. 2004;431:569–73.

    Article  CAS  PubMed  Google Scholar 

  30. Lai J, Li Y, Messing J, Dooner HK. Gene movement by Helitron transposons contributes to the haplotype variability of maize. Proc Natl Acad Sci USA. 2005;102:9068–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Martins C, Ferreira IA, Oliveira C, Foresti F, Galetti Jr PM. A tandemly repetitive centromeric DNA sequence of the fish Hoplias malabaricus (Characiformes: Erythrinidae) is derived from 5S rDNA. Genetica. 2006;127:133.

    Article  CAS  PubMed  Google Scholar 

  32. Cioffi MB, Martins C, Centofante L, Jacobina U, Bertollo LAC. Chromosomal variability among allopatric populations of Erythrinidae fish Hoplias malabaricus: mapping of three classes of repetitive DNAs. Cytogenet Genome Res. 2009;125:132–41.

    Article  CAS  PubMed  Google Scholar 

  33. Senanan W, Kapuscinski AR, Na-Nakorn U, Miller LM. Genetic impacts of hybrid catfish farming (Clarias macrocephalus × C. gariepinus) on native catfish populations in central Thailand. Aquaculture. 2004;235:167–84.

    Article  CAS  Google Scholar 

  34. Na-Nakorn U. Pla Duk”: breeding and farming. Thailand: Reo See Keaw Printing; 1994.

    Google Scholar 

  35. Ingthamjitr S. Hybrid catfish Clarias catfish seed production and marketing in Central Thailand and experimental testing of seed quality. Ph.D. Dissertation. Bangkok: Asian Institute of Technology; 1997.

    Google Scholar 

  36. Campton DE. Natural hybridization and introgression in fishes. In: Ryman N, Utter F, editors. Population genetics and fishery management. Washington Sea Grant Program. Seattle: University of Washington Press; 1987. p. 161–92.

    Google Scholar 

  37. Teugels GG, Guyomard R, Legendre M. Enzymatic variation in African clariid catfishes. J Fish Biol. 1992;40:87–96.

    Article  CAS  Google Scholar 

  38. Porto-Foresti F, Hashimoto DT, Alves AL, Almeida RBC, Senhorini JA, Bortolozzi J, et al. Cytogenetic markers as diagnoses in the identification of the hybrid between piauçu (Leporinus macrocephalus) and piapara (Leporinus elongatus). Genet Mol Biol. 2008;31:195–202.

    Article  CAS  Google Scholar 

  39. Almeida-Toledo LF, Foresti F, Toledo-Filho AS, Bernardino G, Ferrari W, Alcantara RCG. Cytogenetic studies in Colossoma macropomum, Colossoma mitrei and their interespecific hybrids. In: Klaustiews K, editor. Selection, hybridization and genetic engineering in aquaculture vol 1. Berlin: Heenemann Gm BH & Company; 1987. p. 190–5.

    Google Scholar 

  40. Brinn MNA, Porto JIR, Feldberg E. Karyological evidence for interspecific hybridization between Cichla monoculus and C. temensis (Perciformes, Cichlidae) in the Amazon. Hereditas. 2004;141:252–7.

    Article  PubMed  Google Scholar 

  41. Nirchio M, Fenocchio AS, Swarça AC, Pérez JE, Granado A, Estrada A, Ron E: Cytogenetic characterization of hybrids offspring between Colossoma macropomum (Cuvier, 1818) and Piaractus brachypomus (Cuvier, 1817) from Caicara del Orinoco, Venezuela, Caryologia. 2003;56:4, 405-411

  42. Tautz D, Renz M. Simple sequences are ubiquitous repetitive components of eukaryotic genomes. Nucl Acids Res. 1984;12:4127–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lewis WH. Polyploidy in species populations. In: Lewis WH, editor. Polyploidy: biological relevance. New York: Plenum; 1980. p. 103–44.

    Chapter  Google Scholar 

  44. Mallet J. Hybrid speciation. Nature. 2007;446:279–83.

    Article  CAS  PubMed  Google Scholar 

  45. McClintock B. The significance of responses of the genome to challenge. Science. 1984;226:792–801.

    Article  CAS  PubMed  Google Scholar 

  46. Wang J, Ye LH, Liu QZ, Peng LY, Liu W, Yi XG, et al. Rapid genomic DNA changes in allotetraploid fish hybrids. Heredity. 2015;114:601–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Cox MP, Dong T, Shen G, Dalvi Y, Scott DB, Ganley ARD. An interspecific fungal hybrid reveals cross-kingdom rules for allopolyploid gene expression patterns. PLoS Genet. 2014;10:e1004180.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Bertollo LAC, Cioffi MB, Moreira-Filho O. Direct chromosome preparation from Freshwater Teleost Fishes. In: Ozouf-Costaz C, Pisano E, Foresti F, Almeida Toledo LF, editors. Fish cytogenetic techniques (Chondrichthyans and Teleosts). CRC Press: Enfield USA; 2015. p. 21–6.

    Chapter  Google Scholar 

  49. Kubat Z, Hobza R, Vyskot B, Kejnovsky E. Microsatellite accumulation in the Y chromosome of Silene latifolia. Genome. 2008;51:350–56.

    Article  CAS  PubMed  Google Scholar 

  50. Pinkel D, Straume T, Gray J. Cytogenetic analysis using quantitative, high sensitivity, fluorescence hybridization. Proc Natl Acad Sci USA. 1986;83:2934.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Levan A, Fredga K, Sandberg AA. Nomenclature for centromeric position on chromosomes. Hereditas. 1964;52(2):201.

    Article  Google Scholar 

  52. Ifeoluwa OT, Adeogun AO, Bakare AA, Sowumi AA, Agwumba, Ugwumba OA. Karyotypic description of six species of Clarias (Siluriformes: Clariidae) from South West, Nigeria. J Anim Vet Adv. 2011;3(4):264–73.

    Google Scholar 

  53. Wu GM, Xin CB, Chen YL. A comparative study on the karyotypes of four species of catfish. Acta Genet Sinica. 1986;3(3):213–20.

    Google Scholar 

  54. Cui JX, Ren XH, Yu QX. Nuclear DNA content variation in fishes. Cytologia. 1991;56:425–9.

    Article  Google Scholar 

  55. Rishi KK. Chromosomal homogeneity in two Indian catfishes. Geobios. 1978;5:121–3.

    Google Scholar 

  56. Prasad R. Clarias batrachus. In: Ponniah AG, John G, editors. Fish chromosome atlas. National Bureau of Fish Genetic Resources: Lucknow, India; 1998. p. 214–5.

    Google Scholar 

  57. Hinegardner R, Rosen DE. Cellular DNA content and the evolution of teleostean fishes. Amer Natur. 1972;106:621–44.

    Article  CAS  Google Scholar 

  58. Pandey N, Lakra WS. Evidence of female heterogamety, B-chromosome and natural tetraploidy in the Asian catfish, Clarias batrachus, used in aquaculture. Aquaculture. 1997;149:31–7.

    Article  Google Scholar 

  59. Nagpure NS, Kushwah B, Srivastava SK, Ponniah AG. Comparative Karyomorphology of African catfish Clarias gariepinus (Burchell) and Asian catfish Clarias batrachus (Linn.). Chromosome sci. 2000;4:57–9.

    Google Scholar 

  60. Siraj SS, Sukardi R, Phaik Imm CC, Mei KH, Vellasamy S, Panandam JM, et al. Chromosome analysis of walking catfish, Clarias spp. J Malaysian Soc Appl Biol. 2009;38(1):61–4.

    Google Scholar 

  61. Yaqoob S, Tahir Mohi-uddin T, Swaleha. Karyotypic analysis of walking catfish Clarias batrachus (Linnaeus, 1758). Researcher. 2012;4(5):46–8.

    Google Scholar 

  62. Ozouf-Costaz C, Parrent M, Teugels GG, Van der Berg E. Further karyological data on African clariid catfishes (Ostariophysi, Siluroidei). Abstr. VIII Congr. Soc Europ Ichthyol. 1994;1:56.

  63. Luo JL, Wang ZX, Lin SP, Yang JH. Studies on the karyotype of Clarias fuscus. J Fish China. 1986;10:441–6.

    Google Scholar 

  64. Yu XJ, Zhou T, Li YC, Li K, Zhou M. Chromosomes of Chinese freshwater fishes. Beijing: Science Press; 1989.

    Google Scholar 

  65. Arai R, Hirano H. First record of the clariid catfish, Clarias fuscus, from Japan. Japan J Ichthyol. 1974;21:53–60.

    Google Scholar 

  66. Ozouf-costaz C, Teugels GG, Legendre M. Karyological analysis of three strains of the African catfish, Clarias gariepinus (Clariidae), used in Aquaculture. Aquaculture. 1990;87:271–7.

    Article  Google Scholar 

  67. Karahan A, Ergene S. Chromosomal differentiation between populations of Clarias gariepinus (Burchell, 1822) from the Goksu Delta and Orontes River (Turkey). Turk J Biol. 2011;35:79–87.

    Google Scholar 

  68. Omotayo F. Karyological analysis of the African catfish, Clarias gariepinus (Burchell, 1822). Wudpecker J Agric Res. 2012;1(11):674–7.

    Google Scholar 

  69. Awodiran MO, Majolagbe FA, Komolafe OO, Adewumi AA, Oyebola OO, Oyewole RO. Cytogenetic study and serum protein characterization of Clarias gariepinus (Burchell, 1822) and Heterobranchus bidorsalis (Geoffroy Saint-Hillaire, 1809) in South western Nigeria. Int J Biol Chem. 2014;8(6):2377–86.

    Google Scholar 

  70. Donsakul T, Magtoon W. A karyotype of two clariid catfishes, Clarias gariepinus and interspecific hybridization between C. gariepinus and C. macrocephalus. In: The 17th Congress on science and technology of Thailand B-038. 1992. p. 406–7.

    Google Scholar 

Download references

Acknowledgments

This research was funded by the Research Group on Toxic Substances in Livestock and Aquatic Animals, Faculty of Veterinary Medicine and Royal Thai Government scholarship National Science and Technology Development Agency (NSTDA) for financial support and by the Brazilian agencies FAPESP and CNPq.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bundit Tengjaroenkul.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

NM carried out the cytogenetic analysis and drafted the manuscript, MBC helped in the cytogenetic analysis, drafted and revised the manuscript, CFY and NG carried out the molecular cytogenetic analysis, drafted and revised the manuscript, LACB. SD, WS and WFM drafted and revised the manuscript. AT and BT coordinated the study, drafted and revised the manuscript. All authors read and approved the final version of the manuscript.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Maneechot, N., Yano, C.F., Bertollo, L.A.C. et al. Genomic organization of repetitive DNAs highlights chromosomal evolution in the genus Clarias (Clariidae, Siluriformes). Mol Cytogenet 9, 4 (2016). https://doi.org/10.1186/s13039-016-0215-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13039-016-0215-2

Keywords