Next Article in Journal
Screening of New Microsatellite DNA Markers from the Genome of Platyeriocheir formosa
Previous Article in Journal
Effect of Tea Polyphenol on Oxidative Injury in S180 Cells Induced Hepatocarcinoma Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Loss of Genetic Diversity in the Cultured Stocks of the Large Yellow Croaker, Larimichthys crocea, Revealed by Microsatellites

1
State Key Laboratory of Biocontrol, Institute of Aquatic Economic Animals and the Guangdong Province Key Laboratory for Aquatic Economic Animals, School of Life Sciences, Sun Yat-sen University, Guangzhou 510275, China
2
College of Ocean & Earth Sciences, Xiamen University, Xiamen 361005, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2012, 13(5), 5584-5597; https://doi.org/10.3390/ijms13055584
Submission received: 14 March 2012 / Revised: 23 April 2012 / Accepted: 2 May 2012 / Published: 9 May 2012

Abstract

:
The large yellow croaker (Larimichthys crocea) is the most important mariculture fish species in China and the wild stocks of this croaker have collapsed in the past decades due to high fishing pressure and habitat degradation. Due to a lack of wild croaker samples, however, studies concerning the genetic changes of the cultured croaker stocks compared to their wild counterparts were never conducted. Here, we collected three wild populations in the northern and central East China Sea during fisheries survey and investigated the differences in terms of genetic diversity and differentiation between and within cultured stocks and wild populations. Our results demonstrated that the cultured croaker had significantly reduced genetic diversity in contrast to the wild populations, and also presented statistically significant differentiation from the wild, indicating that enhancement of the current wild stock should be conducted with caution. These changes may be caused by founder effects, artificial selection and random genetic drift. With a relatively high level of genetic diversity, the wild populations showed important value for improving the ongoing breeding program of this croaker. Further, we detected no differentiation among the wild populations, suggesting that the wild croaker in the northern and central East China Sea should be considered as one unit for management and conservation.

1. Introduction

The large yellow croaker, Larimichthys crocea, is an economically important marine fish species endemic to China. The production of this croaker reached about 200,000 tons in the mid 1970s and it was once ranked in the top three commercial fish species in mainland China [1,2]. However, the wild resources of this croaker has collapsed in the past decades due to heavy exploitation of spawning and over-wintering aggregations, poor stock management, habitat pollution and climate changes [24]. In order to satisfy the needs of consumers for food and also to protect this species from extinction, the Chinese government has conducted successful artificial mariculture for this croaker since 1985 [5]. The aquaculture production reached approximately 70,000 tons in 2006 [6].
However, aquaculture practices are likely to reduce the genetic diversity and further to cause the loss of disease resistance and environmental adaptability, which greatly limit the potential for selective breeding [79]. Recently, several biological changes including small size and early age of sexual maturation, low growth rates, poor flesh quality and loss of resistance to disease and cold have been identified in the cultured croaker in contrast to the wild populations [2,6]. Such changes were suggested to be associated with overexploitation of the wild stocks and mariculture operations [2,6]. The decline of quality may be caused by loss of genetic diversity in the cultured stocks [8]. Following successful hatchery production, larvae and fingerlings from hatcheries were released by the Chinese government to restore and enhance the wild stocks of this croaker [1012]. Such artificial release and random escape from mariculture stocks into the open marine environment can cause potential harmful effects on the genetic make-up of both wild and reared populations without screening the genetic backgrounds of the two types of stocks [7,13,14]. However, no studies have been performed to monitor such genetic make-up and changes between cultured stocks and wild populations of this croaker because of great difficulties in collecting enough wild samples for population genetic studies.
We successfully collected wild populations of the large yellow croaker in fisheries survey from 2007 to 2010, which allowed us to conduct genetic studies as described above. In addition, microsatellite markers have been successfully used in genetic monitoring of the changes between hatchery stocks and wild populations [1517]. Here, we used 10 microsatellites to analyze the genetic status of both cultured and wild populations of the large yellow croaker in China. The aim of our study was to examine the changes of genetic variation and to assess potential genetic differentiation between cultured and wild populations of this croaker.

2. Results

2.1. Genetic Variation within Populations

Under exact tests, no consistent deviations from HWE in each sample were detected. After sequential Bonferroni correction, only two locus-population pairs (H37-NB and H54-ZJW) were significant (corrected P = 0.0007; Table 1). We also found no evidence of LD among loci. Significant presence of null alleles was detected by Microchecker at one locus, H43. The estimated null allele frequencies in each sample at this locus were more than 0.1 and were statistically significant. As the presence of null alleles may bias the results of genetic variation and particularly differentiation, we therefore excluded this locus from further analysis.
In terms of genetic variation, one locus, H80, was monomorphic in two cultured stocks, DQ and MY (Table 1). The other loci all showed a high level of polymorphism in each sample. The highest A was observed in the wild population YCW (7.899), while the lowest was in cultured stock DQ (4.444). For AR, the highest and lowest values were also found in wild population YCW (7.816) and in cultured stock DQ (4.307), respectively. In terms of heterozygosities including HO and HE, the highest and the lowest values were also identified in wild population and cultured stock, respectively (Table 1). Further investigation revealed a statistically significant reduction of genetic diversity in the cultured stocks compared to the wild populations. In detail, AR varied from 4.307 (DQ) to 5.037 (NB) with a mean of 4.686 in cultured stocks, which was significantly lower than in wild populations ranging from 5.913 (ZJW) to 7.816 (YCW) with a mean of 6.730 (P < 0.01). Similar to AR, HO and HE values were also significantly lower in the cultured stocks (from 0.427 to 0.491 and from 0.462 to 0.517 for HO and HE, respectively) than in the wild populations (from 0.570 to 0.647 and from 0.591 to 0.649 for HO and HE, respectively, P < 0.05). However, FIS values showed no significant differences between the cultured stocks and the wild populations (P > 0.05). The Bottleneck analysis did not detect signals of recent population reduction for each sample whether in test under TPM or in mode-shift test.

2.2. Genetic Differentiation among Populations

The global FST was 0.070 with 95% confidence interval ranging from 0.023 to 0.173 and was highly significant (P < 0.001). Pairwise FST analysis showed that the wild populations were significantly divergent from the cultured stocks after Bonferroni correction for multiple comparisons (Table 2; P < 0.002). Among five cultured stocks, only DQ was significantly different from the others. Interestingly, there was no differentiation detected among the three wild populations (Table 2). Results of AMOVA further supported the significant genetic differentiation between wild populations and cultured stocks, which occupied 10.208% of total genetic variation (P < 0.001; Table 3). FCA analysis also revealed significant differences in terms of allele frequency between the cultured stocks and the wild populations (Figure 1). These results were in accordance with the pairwise FST analysis. In simulations of the Bayesian approach with the program Structure, the mean Ln Likelihood values clearly suggested three clusters as the most likely population structure (Figure 2). The results indicated that almost all the wild individuals were assigned into one cluster, whereas the cultured stocks showed identical genetic properties (Figure 3). However, we observed that the cultured stocks and wild populations were clearly assigned into their own clusters at K = 2 (Figure 3). These results strongly supported the results of the other genetic differentiation studies. In addition, we also observed that the DQ stock was slightly divergent from the other cultured stocks both in FCA and in Structure analysis (Figures 1 and 3), which was consistent with FST results (Table 2).

3. Discussion

3.1. Microsatellites Polymorphism

Investigation of genetic variation of cultured and wild populations of domesticated animals can provide valuable information for breeding programs and also for conservation genetics [8,18]. In this study, we used polymorphic microsatellites to analyze the genetic differences between the cultured stocks and the wild populations of the large yellow croaker. The allele numbers per locus varied from 3.125 to 16.125 and from 0.000 to 0.889, respectively, which is similar to other marine fish species, such as Atlantic cod (Gadus morhua), sea bream (Pagrus major) and the orange-spotted grouper (Epinephelus coioides), suggesting these microsatellites are sufficient to detect genetic variation in the large yellow croaker [1921]. Across the eight samples, however, the average allele number and HE per locus were 5.750 and 0.541, respectively, which is much lower than that (20.6 and 0.79 per locus) found in marine fishes [22]. This result might indicate that the large yellow croaker has reduced in genetic diversity due to high fishing pressure and/or artificial breeding.

3.2. Genetic Variation within Populations

Aquaculture practices have been broadly reported to have the tendency to reduce genetic variability in cultured stocks of fish species [16,17,23]. Aquaculture practices are detrimental to the domestication process of cultured stocks because a high level of genetic variation is related to adaptive fitness and therefore can provide more chances for organisms to survive under the pressure of artificial and natural selection [24]. In China, the cultured individuals of the large yellow croaker were consistently considered to show a picture of small size, early age of sexual maturation, low growth rates, poor flesh quality and loss of resistance to disease and cold, which may be a reflection of genetic diversity loss [2,6]. As expected, in our study, we detected a statistically significant decline of genetic variation in the cultured stocks compared to their wild counterparts of this croaker (Table 1). The genetic variation measures used in our study involved AR, HO and HE. Typically, AR is independent of sample size and is more sensitive to be detected in populations of reduced genetic diversity than heterozygosity, as loss of rare alleles shows little effect on heterozygosity [15,18,25]. Here, we found genetic variation declining not only in terms of AR but also in terms of heterozygosity measures, which likely demonstrates that the reduction of genetic variation in the cultured stocks is not caused by sampling and genotyping bias.
The decline of genetic diversity in cultured stocks is typically considered to be the result of interacting founder effects, random genetic drift, and artificial and natural selection in the cultured environments [8,26,27]. In our study, the reduced genetic variation in croaker stocks could be mainly due to founder effects during the domestication process. The broodstocks of this croaker were initially set up in the mid-1980s, when the wild stock of this species had collapsed [2,6]. In this situation, the broodstocks were mainly from random capture of the wild individuals and therefore consisted of only a small number of individuals. This type of broodstock is prone to have great effects on genetic variation of their offspring and ultimately lead to the loss of genetic diversity in the cultured stocks [27]. In addition, random genetic drift and artificial selection much likely played important roles in leading to the loss of genetic diversity in the cultured stocks. On one hand, as an endangered species, there is no large source population available to supplement the small-sized broodstocks of this croaker. In this case, genetic drift is seldom avoided and the cultured stocks would experience excessive loss of genetic variability [8]. On the other hand, these five cultured stocks have experienced artificial breeding for two to three generations. During such domestication process, artificial selection cannot be avoided in order to obtain fingerlings of high quality. Artificial selection has been broadly reported to possibly reduce genetic diversity [18,26,28]. Apart from these factors, natural selection and mutation may also have important effects on genetic diversity [8], though we cannot clearly differentiate whether the two factors reduce or increase the level of genetic diversity in our work. However, it should be noted that all the cultured samples have not experienced bottleneck. This may be due to the short domestication history of this croaker.

3.3. Genetic Divergence among Populations

With respect to genetic divergence, pairwise FST, AMOVA, FCA and Structure analysis consistently supported the significant differentiation between cultured stocks and wild populations (Tables 2 and 3, Figures 1 and 3). Such significant differentiation between cultured stocks and wild populations was also observed in many other food fish species, such as salmon (Salmo salar), grass carp (Ctenopharyngodon idella) and orange-spotted grouper, which was considered to result from artificial selection, founder effects and genetic drift [15,21,29]. Interestingly, we did not detect the signals of differentiation among the three wild populations (Table 2, Figures 1 and 3), suggesting high gene flow in this croaker. It is common for marine fishes, especially for migratory species to present little divergence because of lacking clear barriers to gene flow within open marine environments [30]. This result provides important information for wild stock management and conservation of this croaker. Among the cultured stocks, we observed that only DQ stock was significantly divergent while the others showed little differentiation from each other (Table 2, Figures 1 and 3). Considering the fact that all stocks except for NB experienced selection, the shallow differentiation among cultured stocks might indicate the effects of artificial selection have not been high enough to be detected.

3.4. Implications for Artificial Breeding and Conservation

In total, we detected statistically significantly less genetic diversity in the cultured stocks of the large yellow croaker in contrast to their wild counterparts, which was mainly caused by founder effects, random genetic drift and artificial selection involved in aquaculture practices. In addition, significant genetic divergence between cultured stocks and wild populations was also observed. Possessing a high level of genetic variation, the wild croaker showed great value for ongoing selective breeding programs by providing more genetic variation. At the same time, due to the lack of differentiation, the wild populations can be considered as one unit for conservation. However, ongoing artificial wild stock enhancement by releasing of cultured croaker fingerlings should be conducted with caution, as previous studies have suggested that release of hatchery fingerlings with low genetic variation would likely reduce the genetic diversity of the wild populations and further lead to loss of adaptation to variable environments for the wild populations [3133]. The stock enhancement programs of this croaker were mainly carried out in the north coast of Zhejiang province and more than ten million cultured fries were released from 2000 to 2009, among which about 60,000 were tagged by hanging scutcheon [34]. Recapture study suggested that the released croaker could not only survive but could also spawn in the following years, although the survival rate was quite low [11,34]. Therefore, it is very likely for the released individuals to be integrated into the wild croaker populations. In our study, we detected that the cultured croaker was less diverse in genetic variability than the wild croaker and was significantly divergent from the wild. If the stocked croaker is incorporated into the wild populations, it would definitely change the genetic make-up of the wild croaker populations and further cause adverse effects on the wild croaker.
Since the 1990s, sample collections of the wild populations of this croaker have rarely been reported. In our study, we collected three samples of the wild croaker in the northern and central East China Sea during fisheries survey for several years, which may suggest that this area is “refugia” of this species. Although the wild stock of this croaker has collapsed, we did not detect the signals of recent bottleneck in the three wild croaker populations using microsatellites. In another study, we analyzed the phylogeography of this croaker using the same wild samples as in this study by sequencing mitochondrial Cytb and COI genes. The results revealed that this species was in the process of population expansion after Pleistocene glaciations, which also suggested few signals of genetic bottleneck in the wild croaker populations [35]. However, previous study has demonstrated that the potential of detecting bottleneck is greatly limited in populations experiencing expansion [36]. Combining the fact that wild croaker was seldom captured, the effective population size of the wild croaker can be rather small. Thus, strict measures by the government, such as reducing fishing pressure and avoiding marine environment pollution in this area must be taken into account to protect this species from extinction.

4. Experimental Section

4.1. Sample Collection

We analyzed five cultured stocks and three wild populations of the large yellow croaker. Among these five cultured stocks, three (DQ, MY and NB) were collected from Ningbo, Zhejiang province, while the other two (XP and SD) were from Ningde, Fujian province. In detail, two (MY and NB) and three stocks (DQ, SD and XP) were the third and second generation offspring of the local fish farms, respectively. The broodstocks of these cultured stocks were founded using brooders collected from the central East China Sea. Due to stock collapse in the wild resources of this croaker, however, it is very difficult to capture the wild fish. Fortunately, we collected three wild populations (ZJW, QDW and YCW) from the northern and central East China Sea during several fisheries surveys between 2007 and 2010. Judging from the body size, all the wild individuals were adult. Detailed information about sample size and location is shown in Figure 4. A small piece of muscle tissue or fin clip was collected and stored in 95% ethanol for DNA extraction.

4.2. Molecular Methods

Genomic DNA was isolated using the standard phenol-chloroform extraction protocol [37]. All sampled individuals were genotyped using ten microsatellite loci, namely, H16, H31, H33, H37, H43, H47, H54, H65, H80 and H82 [38]. For microsatellite genotyping, forward primers were 5′-labeled with a fluorescent dye HEX or 6-FAM. PCR amplification was performed according to the Molecular Ecology Resources Primer Development Consortium [38]. PCR products were separated on an ABI PRISM 3730 DNA automated sequencer (Applied Biosystems) and were measured according to the ROX-500 standard using GeneMapper (Applied Biosystems).
We employed the number of alleles (A), observed heterozygosities (HO), expected heterozygosities (HE) and allele riches (AR) to measure the genetic variation in each sample. These parameters were calculated in FSTAT version 2.9.3.2 [39]. Exact tests of Hardy-Weinberg equilibrium (HWE) in each sample for each locus and linkage disequilibrium (LD) between pairs of loci were tested using the Markov chain methods implemented in Genepop 4.0 [40]. Inbreeding coefficient (FIS) was also estimated using FSTAT version 2.9.3.2 [39]. The genotyping errors and presence of null alleles were checked by using the program Microchecker [41]. The significance of differences in genetic variation between cultured stocks and wild populations were tested using Mann–Whitney U test. Recent population size reduction was examined in the form of heterozygote excess using the program Bottleneck version 1.2.02 [42]. The possibility of recent bottlenecks was tested under two-phase model (TPM, with 90% stepwise-mutation) with 1000 iterations and using the graphical mode-shift test by Luikart et al. [43].
Genetic differentiation among samples was estimated using pairwise Wright’s F-statistics (FST). The significance was tested by a permutation with 10,000 replicates using ARLEQUIN 3.5 [44] with sequential Bonferroni correction at the significance level of 0.05. Analysis of molecular variance (AMOVA) was performed to partition genetic variance hierarchically between the wild populations and cultured stocks using ARLEQUIN 3.5 [44]. We also performed factor correspondence analysis (FCA) using the program GENETIX 4.05 [45] to detect population structure based on allele frequencies. In addition, a Bayesian method was employed to investigate the population structure of all samples using the program Structure 2.2.3 [46]. This program can estimate the number of putative genetic clusters (K) and assign individuals into corresponding clusters. We conducted this analysis under admixture model and ran for 106 iterations with a burn-in length of 106. The most likely K value was inferred by investigating mean Ln likelihood values.

5. Conclusions

In total, our study demonstrated that the cultured croaker had significantly reduced in genetic diversity in contrast to the wild and also presented statistically significant differentiation from the wild, indicating the current wild stock enhancement should be conducted with caution. With a relatively high level of genetic diversity, the wild populations showed important value for the ongoing breeding programs of this croaker. Simultaneously, we detected no differentiation among the wild populations, suggesting that the wild croaker in the northern and central East China Sea should be considered as one unit for management and conservation.

Acknowledgments

This work was supported by the National High Technology R & D Program 863 of China (2009AA09Z401) and the Special Fund for Agro-scientific Research in the Public Interest (200903029), as well as the National Natural Science Foundation of China (grant 40976096).

References

  1. Huang, B.; Walters, C.J. Cohort analysis and population dynamics of large yellow croaker in the China Sea. N. Am. J. Fish. Manag 1983, 3, 295–305. [Google Scholar]
  2. Liu, M.; de Mitcheson, Y.S. Profile of a fishery collapse: Why mariculture failed to save the large yellow croaker. Fish Fish 2008, 9, 219–242. [Google Scholar]
  3. Chen, B.; Gong, Z. Juvenile protection was an essential measure for the recovery of the large yellow croaker stock. J. Fish. Fujian 1984, 1, 17–23. [Google Scholar]
  4. Kong, X.; Hong, G.; Mao, X.; Bi, C. Section Three: Large Yellow Croaker. In Fishery Resource Survey and Area Plan in the East China Sea; East China Normal University Press: Shanghai, China, 1987; pp. 318–338. [Google Scholar]
  5. Breeding and Farming of Pseudosciaena crocea; Su, Y.; Zhang, C.; Wang, J. (Eds.) Ocean Press: Beijing, China, 2007.
  6. Wu, X.W.; Liu, X.D.; Cai, M.Y.; Ye, X.J.; Wang, Z.Y. Genetic analysis of farmed and wild stocks of large yellow croaker Larimichthys crocea by using microsatellite markers. Afr. J. Biotechnol 2011, 10, 5773–5784. [Google Scholar]
  7. Allendorf, F.W.; Phelps, S.R. Loss of genetic variation in a hatchery stock of cutthroat trout. Trans. Am. Fish. Soc 1980, 109, 537–543. [Google Scholar]
  8. Lacy, R.C. Loss of genetic diversity from managed populations: Interacting effects of drift, mutation, selection, and population subdivision. Conserv. Biol 1987, 1, 143–158. [Google Scholar]
  9. Lind, C.E.; Evans, B.S.; Knauer, J.; Taylor, J.J.U.; Jerry, D.R. Decreased genetic diversity and a reduced effective population size in cultured silver-lipped pearl oysters (Pinctada maxima). Aquaculture 2009, 286, 12–19. [Google Scholar]
  10. Liu, J.; Weng, Z.; Tang, X.; Guan, J.; Shi, X.; Hong, G. Tagging techniques of Pseudosciaena crocea (Richardson) in Guanjingyang sea area and a preliminary study on ecological habits and characteristics of early tagged fish. Mar. Sci 1994, 5, 53–58. [Google Scholar]
  11. Xu, H.; Zhou, Y. A preliminary study on release and enhancement of large yellow croaker Pseudosciaena crocea in the north coast of Zhejiang. Mar. Fish 2003, 25, 69–72. [Google Scholar]
  12. Zhou, Y. The retrospection and prospect of releasing and enhancement of fishery resources in Zhejiang coastal waters. Mar. Fish 2004, 26, 131–139. [Google Scholar]
  13. Ryman, N. Conservation genetics considerations in fishery management. J. Fish Biol 1991, 39, 211–214. [Google Scholar]
  14. Ward, R.D. The importance of identifying spatial population structure in restocking and stock enhancement programmes. Fish Res 2006, 80, 9–18. [Google Scholar]
  15. Norris, A.T.; Bradley, D.G.; Cunningham, E.P. Microsatellite genetic variation between and within farmed and wild Atlantic salmon (Salmo salar) populations. Aquaculture 1999, 180, 247–264. [Google Scholar]
  16. Sekino, M.; Hara, M.; Taniguchi, N. Loss of microsatellite and mitochondrial DNA variation in hatchery strains of Japanese flounder Paralichthys olivaceus. Aquaculture 2002, 213, 101–122. [Google Scholar]
  17. Lundrigan, T.A.; Reist, J.D.; Ferguson, M.M. Microsatellite genetic variation within and among Arctic charr (Salvelinus alpinus) from aquaculture and natural populations in North America. Aquaculture 2005, 244, 63–75. [Google Scholar]
  18. Allendorf, F.W. Genetic drift and the loss of alleles versus heterozygosity. Zoo. Biol 1986, 5, 181–190. [Google Scholar]
  19. Bentzen, P.; Taggart, C.T.; Ruzzante, D.E.; Cook, D. Microsatellite polymorphism and the population structure of Atlantic cod (Gadus morhua) in the northwest Atlantic. Can. J. Fish. Aquat. Sci 1996, 53, 2706–2721. [Google Scholar]
  20. Takagi, M.; Taniguchi, N.; Cook, D.; Doyle, R.W. Isolation and characterization of microsatellite loci from red sea bream Pagrus major and detection in closely related species. Fish. Sci 1997, 63, 199–204. [Google Scholar]
  21. Wang, L.; Meng, Z.; Liu, X.; Zhang, Y.; Lin, H. Genetic diversity and differentiation of the orange-spotted grouper (Epinephelus coioides) between and within cultured stocks and wild populations inferred from microsatellite DNA analysis. Int. J. Mol. Sci 2011, 12, 4378–4394. [Google Scholar]
  22. DeWoody, J.A.; Avise, J.C. Microsatellite variation in marine, freshwater and anadromous fishes compared with other animals. J. Fish Biol 2000, 56, 461–473. [Google Scholar]
  23. Kohlmann, K.; Kersten, P.; Flajshans, M. Microsatellite-based genetic variability and differentiation of domesticated, wild and feral common carp (Cyprinus carpio L.) populations. Aquaculture 2005, 247, 253–266. [Google Scholar]
  24. Koljonen, M.L.; Tähtinen, J.; Säisä, M.; Koskiniemi, J. Maintenance of genetic diversity of Atlantic salmon (Salmo salar) by captive breeding programmes and the geographic distribution of microsatellite variation. Aquaculture 2002, 212, 69–92. [Google Scholar]
  25. Comps, B.; Gomory, D.; Letouzey, J.; Thiébaut, B.; Petit, R.J. Diverging trends between heterozygosity and allelic richness during postglacial colonization in the European beech. Genetics 2001, 157, 389–397. [Google Scholar]
  26. Eknath, A.E.; Doyle, R.W. Indirect selection for growth and life-history traits in Indian carp aquaculture: I. Effects of broodstock management. Aquaculture 1985, 49, 73–84. [Google Scholar]
  27. Allendorf, F.W.; Ryman, N. Genetic Management of Hatchery Stocks. Ryman, N., Utter, F., Eds.; University of Washington Press: Seattle, WA, USA, 1987; pp. 141–159. [Google Scholar]
  28. Wright, S.I.; Vroh Bi, I.; Schroeder, S.G.; Schroeder, S.G.; Yamasaki, M.; Doebley, J.F.; McMullen, M.D.; Gaut, B.S. The effects of artificial selection on the maize genome. Science 2005, 308, 1310–1314. [Google Scholar]
  29. Liu, F.; Xia, J.-H.; Bai, Z.-Y.; Fu, J.-J.; Li, J.-L.; Yue, G.H. High genetic diversity and substantial population differentiation in grass carp (Ctenopharyngodon idella) revealed by microsatellite analysis. Aquaculture 2009, 297, 51–56. [Google Scholar]
  30. Waples, R.S. Separating the wheat from the chaff: patterns of genetic differentiation in high gene flow species. J. Hered 1998, 89, 438–450. [Google Scholar]
  31. Cross, T.F.; King, J. Genetic effects of hatchery rearing in Atlantic salmon. Aquaculture 1983, 33, 33–40. [Google Scholar]
  32. Gall, G.A.E. Inbreeding. In Population Genetics and Fishery Management; Ryman, N., Utter, F., Eds.; University of Washington Press: Seattle, WA, USA, 1987; pp. 47–87. [Google Scholar]
  33. Nielsen, J.L.; Gan, C.; Thomas, W.K. Differences in genetic diversity for mitochondrial DNA between hatchery and wild populations of Oncorhynchus. Can. J. Fish. Aquat. Sci 1994, 51, 290–297. [Google Scholar]
  34. Ding, A.; He, Y. Test on release and proliferation of Pseudosciaena crocea in Daiquyang sea area. South China Fish. Sci 2011, 1, 73–77. [Google Scholar]
  35. Wang, L.; Shi, X.; Su, Y.; Meng, Z.; Lin, H. Genetic divergence and historical demography revealed by mtDNA in the large yellow croaker. Larimichthys crocea. Biochem. Syst. Ecol 2012. (submitted for publication). [Google Scholar]
  36. Rogers, A.R.; Harpending, H. Population growth makes waves in the distribution of pairwise genetic differences. Mol. Biol. Evol 1992, 9, 552–569. [Google Scholar]
  37. Sambrook, T.; Fritsch, E.F.; Maniatis, T. Molecular Cloning: A Laboratory Manual, 2nd ed; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 1989. [Google Scholar]
  38. Molecular Ecology Resources Primer Development Consortium. Permanent genetic resources added to Molecular Ecology Resources Database 1 December 2009–31 January 2010. Mol. Ecol. Resour 2010, 10, 576–579.
  39. Goudet, J. Fstat, a program to estimate and test gene diversities and fixation indices. 2002. Available online: http://www2.unil.ch/popgen/softwares/fstat.htm accessed on 5 May 2012.
  40. Rousset, F. Genepop’007: A complete reimplementation of the Genepop software for Windows and Linux. Mol. Ecol. Resour 2008, 8, 103–106. [Google Scholar]
  41. van Oosterhout, C.; Hutchinson, W.F.; Wills, D.P.M.; Shipley, P. Micro-checker: Software for identifying and correcting genotyping errors in microsatellite data. Mol. Ecol. Notes 2004, 4, 535–538. [Google Scholar]
  42. Piry, S.; Luikart, G.; Cournet, J.M. Bottleneck: A computer program for detecting recent reductions in the effective population size using allele frequency data. J. Hered 1999, 90, 502–503. [Google Scholar]
  43. Luikart, G.; Allendorf, F.W.; Cornuet, J.M.; Sherwin, W.B. Distortion of allele frequency distributions provides a test for recent population bottlenecks. J. Hered 1998, 89, 238–247. [Google Scholar]
  44. Excoffier, L.; Laval, G.; Schneider, S. Arlequin ver. 3.0: An integrated software package for population genetics data analysis. Evol. Bioinf. Online 2005, 1, 47–50. [Google Scholar]
  45. GENETIX, version 4.05; Université de Montpellier II: Montpellier, France, 2004.
  46. Pritchard, J.K.; Stephens, M.; Donnelly, P. Inference of stock structure using multilocus genotype data. Genetics 2000, 155, 945–959. [Google Scholar]
Figure 1. Three dimensional scatter plots for individuals of eight samples of the large yellow croaker based on factor correspondence analysis (FCA) analysis, in which each axis represents one principal factor.
Figure 1. Three dimensional scatter plots for individuals of eight samples of the large yellow croaker based on factor correspondence analysis (FCA) analysis, in which each axis represents one principal factor.
Ijms 13 05584f1
Figure 2. The log likelihood over 20 runs for each K values, where the highest value is suggested to be the true number of clusters (K = 3).
Figure 2. The log likelihood over 20 runs for each K values, where the highest value is suggested to be the true number of clusters (K = 3).
Ijms 13 05584f2
Figure 3. Results of structure analysis based on nine microsatellite loci. Each individual is represented by a vertical line, which is colored according to the assigned groups at estimated K = 3 (see Figure 2). Results of K = 2 are also presented for identifying the significant differentiation between cultures stocks and wild populations.
Figure 3. Results of structure analysis based on nine microsatellite loci. Each individual is represented by a vertical line, which is colored according to the assigned groups at estimated K = 3 (see Figure 2). Results of K = 2 are also presented for identifying the significant differentiation between cultures stocks and wild populations.
Ijms 13 05584f3
Figure 4. Maps for sampling localities of the large yellow croaker used in this study. YCW: wild population from the northern East China Sea; QDW and ZJW: wild populations from the central East China Sea. DQ, MY, NB: cultured stocks collected from Ningbo, Zhejiang province; XP and SD: cultured stocks from Ningde, Fujian province. Sample size is showed in number after name abbreviation.
Figure 4. Maps for sampling localities of the large yellow croaker used in this study. YCW: wild population from the northern East China Sea; QDW and ZJW: wild populations from the central East China Sea. DQ, MY, NB: cultured stocks collected from Ningbo, Zhejiang province; XP and SD: cultured stocks from Ningde, Fujian province. Sample size is showed in number after name abbreviation.
Ijms 13 05584f4
Table 1. Summary statistics at nine microsatellite loci across samples of large yellow croaker.
Table 1. Summary statistics at nine microsatellite loci across samples of large yellow croaker.
Locus\SapleDQMYNBSDXPZJWQDWYCW
H16A43343445
AR3.9733.0003.0003.6873.0003.9953.7334.956
FIS0.1360.2760.116−0.329−0.041−0.004−0.2540.054
HO0.5630.4690.5310.8130.5940.6330.7670.609
HE0.6490.6440.6000.6150.5710.6310.6140.643
H31A33433454
AR3.0002.9923.6802.9992.9983.9324.4633.999
FIS−0.136−0.2340.0820.264−0.191−0.208−0.009−0.080
HO0.6880.6560.5000.4060.5630.7670.6000.696
HE0.6070.5340.5440.5500.4740.6370.5950.645
H33A33333344
AR2.9733.0002.9733.0003.0002.9963.4673.913
FIS−0.254−0.183−0.0660.0610.1570.111−0.0590.050
HO0.5000.5310.3440.4060.4380.4670.4000.435
HE0.4000.4500.3230.4320.5180.5240.3780.457
H37A787988811
AR6.9767.0606.6528.7767.6447.5037.65810.825
FIS0.1510.1780.3670.1650.2010.1030.0830.065
HO0.7100.6560.5000.7190.6250.6900.7330.783
HE0.8340.7960.785*0.8590.7790.7680.799
H47A23333344
AR1.9992.9982.9982.8782.6852.4673.6623.912
FIS−0.088−0.0730.3760.376−0.072−0.009−0.074−0.041
HO0.1880.4380.2190.0940.1880.0670.2330.174
HE0.1730.4080.3490.1490.1750.0660.2180.167
H54A1215151513221621
AR11.24613.21213.11713.20611.79019.66914.29121.000
FIS−0.0430.1410.108−0.0540.0720.2670.1600.051
HO0.8440.7810.7810.9380.8130.7000.7670.909
HE0.8100.9070.8740.8900.8750.950*0.910
H65A33422434
AR2.6872.9653.6601.9061.6883.9162.9973.956
FIS0.4630.3910.555−0.0160.000−0.1500.1370.165
HO0.2190.1250.1560.0630.0310.4000.2410.391
HE0.4040.2040.3480.0620.0310.3490.2790.467
H80A113227510
AR1.0001.0002.3751.6881.6886.7014.9689.825
FISNANA−0.0080.0000.0000.1540.175−0.199
HONANA0.0630.0310.0310.6550.5561.000
HENANA0.0620.0310.0310.7730.6720.838
H82A56766788
AR4.9065.9856.8785.9645.9636.9667.9757.955
FIS−0.0790.1960.020−0.101−0.066−0.0280.0300.011
HO0.7810.6250.7500.7810.7500.8330.8330.826
HE0.7250.7750.7650.7110.7040.8110.8590.835
MeanA4.4445.0005.4445.2224.7786.8896.3337.889
AR4.3074.6905.0374.9004.4956.4615.9137.816
FIS0.0240.0940.1760.0110.0310.0550.0370.004
HO0.4910.4680.4270.4720.4480.5790.5700.647
HE0.5030.5160.5170.4780.4620.6120.5910.649
*Significant deviation from Hardy–Weinberg equilibrium (HWE) after Bonferroni correction (P < 0.0007);
NA, not available.
Table 2. Pairwise FST (below diagonal) and associated P values (above diagonal) among cultured stocks and wild populations of the large yellow croaker. Significant P values after Bonferroni correction for multiple comparisons are denoted in bold.
Table 2. Pairwise FST (below diagonal) and associated P values (above diagonal) among cultured stocks and wild populations of the large yellow croaker. Significant P values after Bonferroni correction for multiple comparisons are denoted in bold.
SamplesDQMYNBSDXPZJWQDWYCW
DQ00.0000.0010.0000.0000.0000.0000.000
MY0.03200.2480.0080.0010.0000.0000.000
NB0.0250.00500.0340.0390.0000.0000.000
SD0.0320.0160.01200.0950.0000.0000.000
XP0.0520.0210.0120.00700.0000.0000.000
ZJW0.1190.1170.1270.1340.14300.0480.011
QDW0.1070.1040.1050.1210.1380.00900.049
YCW0.0960.0960.0920.1000.1120.0150.0100
Table 3. Results of analysis of molecular variance (AMOVA) between cultured stocks and wild populations of the large yellow croaker.
Table 3. Results of analysis of molecular variance (AMOVA) between cultured stocks and wild populations of the large yellow croaker.
Source of VariationSum of SquaresVariance ComponentsPercentage VariationP Value
Among groups64.6480.28010.2080.000
Among populations within groups31.0720.0431.5760.000
Among individuals within populations597.8070.1334.8530.000
Within individuals554.0002.28783.3630.000
Total1247.5262.743100.000

Share and Cite

MDPI and ACS Style

Wang, L.; Shi, X.; Su, Y.; Meng, Z.; Lin, H. Loss of Genetic Diversity in the Cultured Stocks of the Large Yellow Croaker, Larimichthys crocea, Revealed by Microsatellites. Int. J. Mol. Sci. 2012, 13, 5584-5597. https://doi.org/10.3390/ijms13055584

AMA Style

Wang L, Shi X, Su Y, Meng Z, Lin H. Loss of Genetic Diversity in the Cultured Stocks of the Large Yellow Croaker, Larimichthys crocea, Revealed by Microsatellites. International Journal of Molecular Sciences. 2012; 13(5):5584-5597. https://doi.org/10.3390/ijms13055584

Chicago/Turabian Style

Wang, Le, Xiaofeng Shi, Yongquan Su, Zining Meng, and Haoran Lin. 2012. "Loss of Genetic Diversity in the Cultured Stocks of the Large Yellow Croaker, Larimichthys crocea, Revealed by Microsatellites" International Journal of Molecular Sciences 13, no. 5: 5584-5597. https://doi.org/10.3390/ijms13055584

Article Metrics

Back to TopTop