Hostname: page-component-848d4c4894-r5zm4 Total loading time: 0 Render date: 2024-06-21T07:09:36.920Z Has data issue: false hasContentIssue false

Sympatric morphological and genetic differentiation of the pearl oyster Pinctada radiata (Bivalvia: Pterioida) in the northern Persian Gulf

Published online by Cambridge University Press:  04 November 2014

Moein Rajaei
Affiliation:
Department of Fisheries, Faculty of Natural Resources, University of Tehran, Karaj PO Box 4314, Iran
Hamid Farahmand*
Affiliation:
Department of Fisheries, Faculty of Natural Resources, University of Tehran, Karaj PO Box 4314, Iran
Hadi Poorbagher
Affiliation:
Department of Fisheries, Faculty of Natural Resources, University of Tehran, Karaj PO Box 4314, Iran
Mohammad Sedigh Mortazavi
Affiliation:
The Persian Gulf and Oman Sea Ecology Research Centre, Bandar Abbas, Iran
Ahmad Farhadi
Affiliation:
Department of Fisheries, Faculty of Natural Resources, University of Tehran, Karaj PO Box 4314, Iran
*
Correspondence should be addressed to:H. FarahmandDepartment of Fisheries, Faculty of Natural Resources, University of Tehran, Karaj, PO Box 4314, Iran email: hfarahmand@ut.ac.ir

Abstract

The pearl oyster, Pinctada radiata, shows great variation in shell morphology throughout its distribution. This variation can be related to phenotypic plasticity, genetic variability or a combination of both. Using geometric morphometric and microsatellite DNA analyses, two morphologically distinct populations of the pearl oyster were studied in the northern Persian Gulf, i.e. from the Lavan and Hendourabi Islands. Ten landmarks were selected to define the shape of the left shell. In addition, concentration of Zn, Mg, Fe, Cu, Pb, Cd, Mn and Cr of the soft tissues were measured using atomic absorption spectrometry. Six microsatellite loci were used to assess the population genetic structure of the pearl oyster. There were morphometric differences between the populations suggesting the existence of two morphotypes. There was a significant difference between the two populations in concentrations of Fe, Mg, Zn, Cd, Mn and Cr indicating that the specimens from the Lavan Island experience a more stressful environment than those from the Hendourabi Island. Analysis of molecular variance (AMOVA) indicated that the proportion of the genetic variation attributed to differences among populations of the pearl oyster was highly significant for both FST and RST (FST = 0.066, RST = 0.265, P < 0.001). Our findings showed that stressful conditions resulting from heavy metals may have a direct influence on the separation of the populations in Lavan and Hendourabi despite the lack of a physical barrier.

Type
Research Article
Copyright
Copyright © Marine Biological Association of the United Kingdom 2014 

Access options

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

References

REFERENCES

Amos, W. and Harwood, J. (1998) Factors affecting levels of genetic diversity in natural populations. Philosophical Transactions of the Royal Society B: Biological Sciences 353, 177186.CrossRefGoogle ScholarPubMed
Andersen, V., Maage, A. and Johannessen, P. (1996) Heavy metals in blue mussels (Mytilus edulis) in the Bergen Harbor area, western Norway. Bulletin of Environmental Contamination and Toxicology 57, 589596.Google Scholar
Beaumont, A.R. and Khamdan, S.A.A. (1991) Electrophoretic and morphometric characters in population differentiation of the pearl oyster, Pinctada radiata (Leach), from around Bahrain. Journal of Molluscan Studies 57, 433441.Google Scholar
Benzie, J. and Smith-Keune, C. (2006) Microsatellite variation in Australian and Indonesian pearl oyster Pinctada maxima populations. Marine Ecology Progress Series 314, 197211.Google Scholar
Elisabeth, S. (2008) Introduction. In Southgate, P. and Lucas, J. (eds) The pearl oyster. Oxford: Elsevier, pp. 135.Google Scholar
Evans, B.S., Knauer, J., Taylor, J.J.U. and Jerry, D.R. (2006) Development and characterization of six new microsatellite markers for the silver- or gold-lipped pearl oyster, Pinctada maxima (Pteriidae). Molecular Ecology Notes 6, 835837.Google Scholar
Excoffier, L. and Lischer, H.E. (2010) Arlequin suite ver 3.5: a new series of programs to perform population genetics analyses under Linux and Windows. Molecular Ecology Resources 10, 564567.Google Scholar
Excoffier, L., Smouse, P.E. and Quattro, J.M. (1992) Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics 131, 479491.Google Scholar
Garza, J.C. and Williamson, E. G. (2001) Detection of reduction in population size using data from microsatellite loci. Molecular Ecology 10, 305318.CrossRefGoogle ScholarPubMed
Gervis, M.H. and Sims, N.A. (1992) The biology and culture of pearl oysters (Bivalvia pteriidae). London: Overseas Development Administration of the United Kingdom.Google Scholar
Gifford, S.P., MacFarlane, G.R., O'Connor, W.A. and Dunstan, R.H. (2006) Effect of the pollutants lead, zinc, hexadecane and octocosane on total growth and shell growth in the Akoya pearl oyster, Pinctada imbricata. Journal of Shellfish Research 25, 159165.CrossRefGoogle Scholar
Gopurenko, D. and Hughes, J.M. (2002) Regional patterns of genetic structure among Australian populations of the mud crab, Scylla serrata (Crustacea: Decapoda): evidence from mitochondrial DNA. Marine and Freshwater Research 53, 849857.Google Scholar
Goudet, J. (1995) FSTAT (version 1.2): a computer program to calculate F-statistics. Journal of Heredity 86, 485486.Google Scholar
Hoffman, J., Peck, L., Hillyard, G., Zieritz, A. and Clark, M. (2010) No evidence for genetic differentiation between Antarctic limpet Nacella concinna morphotypes. Marine Biology 157, 765778.Google Scholar
Kelly, S.A., Panhuis, T.M. and Stoehr, A.M. (2012) Phenotypic plasticity: molecular mechanisms and adaptive significance. Comprehensive Physiology 2, 14171439.Google Scholar
Kuang, Y., Tong, G., Yan, X. and Sun, X. (2009) Rapid isolation and characterization of microsatellites from the genome of pearl oyster (Pinctada martensi Dunker). Conservation Genetics 10, 14631467.Google Scholar
Labate, J.A. (2000) Software for population genetic analyses of molecular marker data. Crop Science 40, 15211528.CrossRefGoogle Scholar
Leberg, P. (2008) Estimating allelic richness: effects of sample size and bottlenecks. Molecular Ecology 11, 24452449.Google Scholar
Lind, C.E., Evans, B.S., Taylor, J.J.U. and Jerry, D.R. (2007) Population genetics of a marine bivalve, Pinctada maxima, throughout the Indo-Australian Archipelago shows differentiation and decreased diversity at range limits. Molecular Ecology 16, 51935203.Google Scholar
Nei, M. (1973) Analysis of gene diversity in subdivided populations. Proceedings of the National Academy of Sciences 70, 33213323.Google Scholar
Peakall, R. and Smouse, P.E. (2005) GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Notes 6, 288295.Google Scholar
PGSC (2013) Persian Gulf Study Center, Institute of Geopolitical, Strategic, Historical, Geographical studies of the Persian Gulf. Available at http://www.persiangulfstudies.com.Google Scholar
Rajaei, M., Poorbagher, H., Farahmand, H., Mortazavi, M.S. and Eagderi, S. (2014) Inter-population differences in shell forms of the pearl oyster, Pinctada imbricata radiata (Bivalvia: Pterioida) in northern Persian Gulf inferred from principal components analysis (PCA) and elliptic Fourier analysis (EFA). Turkish Journal of Zoology 38, 4248.Google Scholar
Roesijadi, G. and Robinson, W. (1994) Metal regulation in aquatic animals: mechanisms of uptake, accumulation, and release. In Malins, D.C. and Ostrander, G.K. (eds) Aquatic toxicology: molecular, biochemical, and cellular perspective. Boca Raton, FL: Lewis Publishers, pp. 387420.Google Scholar
Rohlf, F. (2009) tpsDig. Version 2.14. Department of Ecology and Evolution, State University of New York.Google Scholar
Rohlf, F. (2010) tpsUtil program, Version 1.46. Department of Ecology & Evolution, State University of New York.Google Scholar
Sambrook, J., Fritsch, E. and Maniatis, T. (1989) Molecular cloning: a laboratory manual. 2nd edn.Cold Spring Harbor, NY: Cold Spring Harbor Laboratory.Google Scholar
Sheets, H. (2001) CoordGen6f, CVAGen6m, PCAGen6n, Regress6k, Tmorphgen6a disponible en IMP. Buffalo, NY: Dept. of Physics, Canisius College; Buffalo, NY: Dept. of Geology, SUNY at Buffalo.Google Scholar
Shi, Y., Wang, Y., Hong, K., Hou, Z., Wang, A. and Guo, X. (2009) Characterization of 31 EST-derived microsatellite markers for the pearl oyster Pinctada martensii (Dunker). Molecular Ecology Resources 9, 177179.Google Scholar
Smith, C., Benzie, J.A.H. and Wilson, K.J. (2003) Isolation and characterization of eight microsatellite loci from silver-lipped pearl oyster Pinctada maxima. Molecular Ecology Notes 3, 125127.Google Scholar
Tlig-Zouari, S., Rabaoui, L., Irathni, I., Diawara, M. and Ben Hassine, O. (2010) Comparative morphometric study of the invasive pearl oyster Pinctada radiata along the Tunisian coastline. Biologia 65, 294300.Google Scholar
Wada, K. (1982) Inter-and intraspecific electrophoretic variation in three species of the pearl oysters from the Nansei Islands of Japan. Bulletin of National Research Institute of Aquaculture 3, 110.Google Scholar
Ward, R.D., Ovenden, J.R., Meadows, J.R., Grewe, P.M. and Lehnert, S.A. (2006) Population genetic structure of the brown tiger prawn, Penaeus esculentus, in tropical northern Australia. Marine Biology 148, 599607.Google Scholar
Weir, B.S. and Cockerham, C.C. (1984) Estimating F-statistics for the analysis of population structure. Evolution 38, 13581370.Google Scholar