首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 11 毫秒
1.
To determine the primary structure of the α-amylase produced by Bacillus subtilis var. amylosacchariticus, we have reported the isolation of thirty-four tryptic peptides and eight CNBr fragments from the enzyme. Since the alignment of the eight CNBr fragments was made by matching with six methionine-containing tryptic peptides, the order of tryptic peptides within each CNBr fragment was determined. In the case of four small CNBr fragments, sequence analyses using an automated sequence analyzer established the peptide orders within these fragments. For larger fragments, further fragmentation was done using chymotrypsin or staphylococcal protease V8 and the resultant peptides were isolated and sequenced. Consequently, the peptide orders within three out of four large CNBr fragments were established.  相似文献   

2.
Stands of canyon live oak (Quercus chrysolepis, Fagaceae) are maintained for fuelwood, fire management, recreation, and as habitat for wildlife. Information about the link between the oak's reproductive ecology and its extent of genetic diversity is important in developing land management policies that will maintain the long-term viability of populations. Basal sprouting is the primary means of reproduction following fire or cutting, and stands frequently include groups of visibly connected trees in a clustered distribution that suggests cloning. We determined the extent to which clusters of trees were clonal and defined the spatial pattern and diversity of genotypes for six populations across nearly the entire east-west extent of the San Bernardino Mountains in southern California. We mapped over 100 trees at each of five sites and genotyped each tree for allozymes at seven polymorphic loci. We identified clones using these multilocus genotypes and detected an average of 34.4 ± 7.3 (SD) clones per site, most of which had unique genotypes. In general, clustered trees belong to single clones and most clones consist of few trees (mean = 3.4 ± 0.6 trees per clone). However, clone size increased significantly with increased individual heterozygosity, suggesting that selection may favor highly heterozygous clones. Clonal diversity and evenness were high relative to reports for most other clonal species; an average of 97% of clones had distinct genotypes, and Simpson's index of diversity averaged 0.95 ± 0.02. Population genetic analyses of 319 clones from six sites revealed high genetic diversity within sites (mean HS = 0.443). Only a small proportion of the total genetic diversity was explained by variation among sites (mean GST = 0.018), which is consistent with high gene flow among sites (Nm = 9.5). We found no significant substructure among plots within sites, and fixation indices within sites were generally small, suggesting that either little inbreeding occurs, and/or few inbred progeny survive. However, spatial autocorrelation analysis of clones indicated fine-scale genetic structure at distances under 4 m, possibly due to limited seed dispersal. Our data suggest that guidelines for seed collection of canyon live oak for use in restoration can be specified in a manner similar to that recommended for conifer species within the region studied.  相似文献   

3.
The population structure of the black rockfish, Sebastes inermis (Sebastidae), was estimated using 10 microsatellite loci developed for S. schlegeli on samples of 174 individuals collected from three wild and three hatchery populations in Korea. Reduced genetic variation was detected in hatchery strains [overall number of alleles (N(A)) = 8.07; allelic richness (A(R)) = 7.37; observed heterozygosity (H(O)) = 0.641] compared with the wild samples (overall N(A) = 8.43; A(R) = 7.83; H(O) = 0.670), but the difference was not significant. Genetic differentiation among the populations was significant (overall F(ST) = 0.0237, P < 0.05). Pairwise F(ST) tests, neighbor-joining tree, and principal component analyses showed significant genetic heterogeneity among the hatchery strains and between wild and hatchery strains, but not among the wild populations, indicating high levels of gene flow along the southern coast of Korea, even though the black rockfish is a benthic, non-migratory marine species. Genetic differentiation among the hatchery strains could reflect genetic drift due to intensive breeding practices. Thus, in the interests of optimal resource management, genetic variation should be monitored and inbreeding controlled within stocks in commercial breeding programs. Information on genetic population structure based on cross-species microsatellite markers can aid in the proper management of S. inermis populations.  相似文献   

4.
Nine Chinese yak breeds (Maiwa,Tianzhu White,Qinghai Plateau,Sibu,Zhongdian,Pall,Tibetan High Mountain,Jiulong,and Xin-jiang) and Gayal were analyzed by means of 16 microsatellite markers to determine the level of genetic variation within populations,genetic relationship between populations,and population structure for each breed.A total of 206 microsatellite alleles were observed.Mean F-statistics (0.056) for 9 yak breeds indicated that 94.4% of the genetic variation was observed within yak breeds and 5.6% of the genetic variation existed amongst breeds.The Neighbor-Joining phylogenetic free was constructed based on Nei's standard genetic dis-tances and two clusters were obtained.The Gayal separated from the yaks far away and formed one cluster and 9 yak breeds were grouped together.The analysis of population structure for 9 yak breeds and the Gayal showed that they resulted in four clusters; one clus-ter includes yaks from Tibet Autonomous Region and Qinghai Province,one cluster combines Zhongdian,Maiwa,and Tianzhu White,and Jiulong and Xinjiang come into the third cluster.Pali was mainly in the first cluster (90%),Jiulong was mainly in the second cluster (87.1%),Zhongdian was primarily in the third cluster (83%),and the other yak breeds were distributed in two to three clusters.The Gayal was positively left in the fourth cluster (99.3%).  相似文献   

5.
We aimed to characterize the population genetic structure within and among five Bemisia tabaci (Gennadius) populations collected from different host plants and geographic regions by using microssatelites as a molecular marker. Each population was represented by 19 specimens. The host plants and geographic origins of these populations were described as follows: Pop 1: Squash Barreiras (BA); Pop 2: Cotton Barreiras (BA); Pop 3: Soybean Campinas (SP); Pop 4: Tomato Cruz das Almas (BA); and Pop 5: Soybean Rondonópolis (MT). Six polymorphic loci were observed, which discriminated 31 different alleles in the studied populations, with a mean number of alleles per population of 3.30 (2.67 - 4.00). Using Fisher's Exact test, it was observed that at least three populations were in Hardy-Weinberg equilibrium for most of the studied loci (six). The dendrogram (UPGMA) separated populations into groups mainly related to the geographic origin of the samples. Only population 5 differed from the others at a 0.15 distance (74.5% group consistency). The most similar populations were 1 and 2, with a 0.01 distance (65.3%). This is in agreement with their geographic origins and it was not consistent with host specificity. The results suggest considerable gene flow (7.3%) among all whitefly populations and indicate that a better understanding of the gene flow in populations of B. tabaci associated with different hosts is required for the management of this insect.  相似文献   

6.
7.
Genetic variability in ten populations of wild-growing ginseng was assessed using AFLP markers with the application of fragment analysis on a genetic analyzer. The variation indices were high in the populations (P = 55.68%; H(S) = 0.1891) and for the species (P = 99.65%; H(S) = 0.2857). Considerable and statistically significant population differentiation was demonstrated (theta = 0.363; Bayesian approach, "full model"; F(ST) = 0.36, AMOVA). The results of AMOVA and Bayesian analysis indicate that 64.46% of variability is found within the populations. Mantel test showed no correlation between the genetic and geographic distances among the populations (r = -0.174; P = 0.817). Hierarchical AMOVA and analysis of genetic relationships based on Euclidean distances (NJ, PCoA, and MST) identified two divergent population groups of ginseng. Low gene flow between these groups (N(m) = 0.4) suggests their demographic independence. In accordance to the concept of evolutionary significant units (ESU), these population groups, in terms of the strategy and tactics for conservation and management of natural resources, should be treated as management units (MUs). The MS tree topology suggests recolonization of southern Sikhote-Alin by ginseng along two directions, from south and west.  相似文献   

8.
Information about the genetic diversity and population structure in elite breeding material is of fundamental importance for the improvement of crops. The objectives of our study were to (a) examine the population structure and the genetic diversity in elite maize germplasm based on simple sequence repeat (SSR) markers, (b) compare these results with those obtained from single nucleotide polymorphism (SNP) markers, and (c) compare the coancestry coefficient calculated from pedigree records with genetic distance estimates calculated from SSR and SNP markers. Our study was based on 1,537 elite maize inbred lines genotyped with 359 SSR and 8,244 SNP markers. The average number of alleles per locus, of group specific alleles, and the gene diversity (D) were higher for SSRs than for SNPs. Modified Roger’s distance (MRD) estimates and membership probabilities of the STRUCTURE matrices were higher for SSR than for SNP markers but the germplasm organization in four heterotic pools was consistent with STRUCTURE results based on SSRs and SNPs. MRD estimates calculated for the two marker systems were highly correlated (0.87). Our results suggested that the same conclusions regarding the structure and the diversity of heterotic pools could be drawn from both markers types. Furthermore, although our results suggested that the ratio of the number of SSRs and SNPs required to obtain MRD or D estimates with similar precision is not constant across the various precision levels, we propose that between 7 and 11 times more SNPs than SSRs should be used for analyzing population structure and genetic diversity.  相似文献   

9.
Major habitats for the snow crab Chionoecetes opilio are mostly found within the northwest Atlantic and North Pacific Oceans. However, the East Sea populations of C. opilio, along with its relative the red snow crab (C. japonicas), are two of the most important commercial crustacean species for fisheries on the east coast of the Korean Peninsula. The East Sea populations of C. opilio are facing declining resources due to overfishing and global climate change. Thus, an analysis of population structure is necessary for future management. Five Korean and one Russian group of C. opilio were analyzed using nine microsatellite markers that were recently developed using next-generation sequencing. No linkage disequilibrium was found between any pair of loci, indicating that the markers were independent. The number of alleles per locus varied from 4 to 18 with a mean of 12, and allelic richness per locus ranged from 4.0 to 17.1 across all populations with a mean of 9.7. The Hardy–Weinberg equilibrium test revealed significant deviation in three out of nine loci in some populations after sequential Bonferroni correction and all of them had higher expected heterozygosity than observed heterozygosity. Null alleles were presumed in four loci, which explained the homozygosity in three loci. The pairwise fixation index (F ST ) values among the five Korean snow crab populations did not differ significantly, but all of the pairwise F ST values between each of the Korean snow crab populations and the Russian snow crab population differed significantly. An UPGMA dendrogram revealed clear separation of the Russian snow crab population from the Korean snow crab populations. Assignment tests based on the allele distribution discriminated between Korean and Russian origins with 93 % accuracy. Therefore, the snow crab populations around the Korean Peninsula need to be managed separately from the populations in Bering Sea in global scale resource management. Also, this information can be used for identification of snow crab origin which is problematic in worldwide crab trade.  相似文献   

10.
In the Mediterranean area, the production of persimmon (Diospyros kaki Thumb) [2n = 6x = 90] has increased recently as an alternative to the major fruit crops. In Spain, production relies almost exclusively on the cultivar “Rojo Brillante” which accounts for 83% of the crop. A crop based on a monovarietal culture implies several commercial risks that can compromise the future of the crop. Although the species was introduced in Europe very recently, it is well adapted to the climate of southern Europe. However, the recent introduction from Japan, the mistakes on the identity of varieties in the collections due to a bad translation of variety names from Japanese, and the lack of genetic characterization of many varieties have caused difficulties for effective management of the available genetic resources. The present paper was aimed at exploring the genetic diversity among different persimmon cultivars, including those collected in the European survey as well as Japanese cultivars. Seventy-one persimmon cultivars coming from two European collections that included accessions from Japan, Italy, and Spain were analyzed using 19 polymorphic microsatellite markers. A total of 206 alleles were obtained, with a mean value of 10.8 alleles per locus. A neighbor joining dendrogram and a principal coordinate analysis arranged the cultivars according to their genetic relationships. Analysis of molecular variance revealed significant genetic variability between and within groups, 73.3% and 85.2% for astringent-type and country origin, respectively. The simple sequence repeat markers classified the persimmon cultivars according to their genetic relationship.  相似文献   

11.
Sixteen polymorphic microsatellite (SSR) markers, developed from an SSR-enriched genomic DNA library of sesame (Sesamum indicum L.), were used to assess genetic diversity, phylogenetic relationships, and population structure among 150 sesame accessions collected from 22 countries. A total of 121 alleles were detected among the sesame accessions. The number of detected alleles varied from 2 to 18, with an average of 7.6 alleles per locus. Polymorphism information content values ranged from 0.03 to 0.79, with an average of 0.42. These values indicated an excess of heterozygous individuals at 16 loci and an excess of homozygous individuals at three loci. Of these, 32 genotype-specific alleles were identified at 11 of 16 polymorphic SSR markers. Cluster analyses were performed by accession and population, revealing a complex accession distribution pattern with mean genetic similarity coefficient of 0.45 by accession and 0.52 by population. The wide variation in genetic similarity among the accessions revealed by SSRs reflected a high level of polymorphism at the DNA level. Model-based structure analysis revealed the presence of three groups that were basically consistent with the clustering results based on genetic distance. These findings may be used to augment the sesame germplasm and to increase the effectiveness of sesame breeding.  相似文献   

12.
The genetic variability for a sample of 227 animals from three populations of Pantaneiro horses was estimated using data from 10 microsatellite loci. The number of alleles and the proportion of heterozygosity indicated high variability. A total of 91 alleles were found, with a significantly high mean number of alleles. The mean polymorphic information content was 0.7 and the paternity exclusion probability was 99.3%. The inbreeding coefficient (F(IS)) was low for the three populations: Ipiranga (F(IS) = 0.147), Nova Esperan?a (F(IS) = 0.094) and Promiss?o (F(IS) = 0.108). Genetic differentiation among all three populations was low (F(ST) = 0.008 to 0.064). Three methods were used to test for a recent bottleneck effect. The graphical method and the Wilcoxon test using the stepwise mutation model showed no bottleneck pattern for any of the populations. The test by two-phase mutation model showed genetic signatures of bottleneck for Ipiranga and Promiss?o. When we consider standard deviation value for Nova Esperan?a, the M-statistic detected a bottleneck pattern, but this result could be explained by a sample size effect. Therefore, there is no immediate cause for concern regarding loss of variation within the breed.  相似文献   

13.
Modern genomics approaches rely on the availability of high-throughput and high-density genotyping platforms. A major breakthrough in wheat genotyping was the development of an SNP array. In this study, we used a diverse panel of 172 elite European winter wheat lines to evaluate the utility of the SNP array for genomic analyses in wheat germplasm derived from breeding programs. We investigated population structure and genetic relatedness and found that the results obtained with SNP and SSR markers differ. This suggests that additional research is required to determine the optimum approach for the investigation of population structure and kinship. Our analysis of linkage disequilibrium (LD) showed that LD decays within approximately 5–10 cM. Moreover, we found that LD is variable along chromosomes. Our results suggest that the number of SNPs needs to be increased further to obtain a higher coverage of the chromosomes. Taken together, SNPs can be a valuable tool for genomics approaches and for a knowledge-based improvement of wheat.  相似文献   

14.
The name Lithocarpus longinux (Hu) Chun ex Y. C. Hsu & H. W. Jen is reinstated. It used to be treated as a synonym of L. areca (Hickel & A. Camus) A. Camus, but morphological characters of cupules and leaves support the reinstatement. These two species, together with L. longzhouicus (C. C. Huang & Y. T. Chang) J. Q. Li & L. Chen, make up a small group that is distributed in limestone areas and have similar morphological characters. A key is provided to distinguish between them, and their is discussed.  相似文献   

15.
A set of 493 old and local Spanish accessions of apple (Malus x domestica Borkh) maintained at three collections in Northeastern Spain was studied using 16 simple sequence repeats in order to estimate their genetic diversity and to identify the genetic structure and relationships among their accessions. An additional diverse set of 45 apple cultivars, including old Spanish and international cultivars, was added as reference. Genetic analyses performed by Bayesian model-based clustering revealed a very strong differentiation of two major groups. The first one clustered 159 individuals (52?% of unique genotypes) including local accessions and six old Spanish cultivars. The second major group was formed by 145 individuals, including 38 international reference cultivars and one old Spanish cultivar. Nested Bayesian clustering was applied to those two groups and two and four sub-groups were found at each one, respectively. The identification of private and unique alleles, and the remarkable differences in allelic richness among groups and sub-groups constitute further evidence of a clear genetic structure. The results obtained through the factorial correspondence and analyses of molecular variance confirmed those obtained by Bayesian analyses, revealing moderate but significant differentiation among the two major groups (F ST?=?0.076) and the six sub-groups (F ST?=?0.111). Our results highlight that the genetic diversity encompassed by currently cultivated apple accounts only for a small fraction of that existing within the species, and that an important part (??60?%) of the local material analyzed constitutes a good example of genetic distinctness with respect to the main cultivars used in European orchards.  相似文献   

16.
We determined the genetic diversity and evolutionary relationships among 26 Chinese indigenous horse breeds and two introduced horse breeds by genotyping these animals for 27 microsatellite loci. The 26 Chinese horse breeds come from 12 different provinces. Two introduced horse breeds were the Mongolia B Horse from Mongolia and the Thoroughbred Horse from the UK. A total of 330 alleles were detected, and the expected heterozygosity ranged from 0.719 (Elenchuns) to 0.780 (Dali). The mean number of alleles among the horse breeds ranged from 6.74 (Hequ) to 8.81 (Debao). Although there were abundant genetic variations found, the genetic differentiation was low between the Chinese horses, which displayed only 2.4% of the total genetic variance among the different breeds. However, genetic differentiation (pairwise FST) among Chinese horses, although moderate, was still apparent and varied from 0.001 for the Guizou–Luoping pair to 0.064 for the Jingjiang–Elenchuns pair. The genetic differentiation patterns and genetic relationships among Chinese horse breeds were also consistent with their geographical distribution. The Thoroughbred and Mongolia B breeds could be discerned as two distinct breeds, but the Mongolia B horse in particular suffered genetic admixture with Chinese horses. The Chinese breeds could be divided into five major groups, i.e. the south or along the Yangtze river group (Bose, Debao, Wenshan, Lichuan, Jianchang, Guizhou, Luoping, Jinjiang and Dali), the Qinghai‐Tibet Plateau group (Chaidamu, Hequ, Datong, Yushu, Tibet Grassland and Tibet Valley), the Northeast of China group (Elenchuns, Jilin and Heihe), the Northwest of China group (Kazakh, Yili and Yanqi) and the Inner Mongolia group (Mongolia A, Sanhe, Xinihe,Wuzhumuqin and Sengeng). This grouping pattern was further supported by principal component analysis and structure analysis.  相似文献   

17.
Genotype data from 20 microsatellites typed in 253 animals is used here to assess the genetic structure of seven European pedigree cattle breeds. Estimation of genetic subdivision using classical drift-based measures shows that the average proportion of genetic variation among breeds varies between 10 and 11% of the total, depending on the estimator used. We demonstrate that a simple allele-sharing genetic distance parameter can be used to construct a dendrogram of relationships among animals. This phylogenetic tree displays a remarkable degree of breed clustering and reflects an extensive underlying kinship structure, particularly for the Swiss Simmental breed and four breeds originating from the British Isles. Condensation of allele frequencies and individual genotypic compositions using principal component analysis is also used to investigate genetic structure among breeds and individual animals. In addition, the underlying genetic demarcation of European cattle breeds is emphasized in simulations of breed assignment using allele frequency distributions from samples of microsatellite loci. Correct breed designation can be inferred with accuracies approaching 100% using data from a panel of 10 microsatellite loci.  相似文献   

18.
? Premise of the study: We isolated and characterized polymorphic microsatellite loci in Cyclobalanopsis glauca (Fagaceae), an evergreen broadleaved monoecious tree, to provide tools for analyzing parentage and mating system. ? Methods and Results: Thirteen polymorphic microsatellite markers were developed and tested in three C. glauca populations. The number of alleles per locus varied from two to 22. The observed and expected heterozygosities within populations were 0.000-0.967 and 0.033-0.949, respectively. ? Conclusions: These polymorphic primers showed high levels of polymorphism within tested populations, and can be used in parentage analysis and mating system estimation of C. glauca.  相似文献   

19.
20.
Ball AO  Chapman RW 《Molecular ecology》2003,12(9):2319-2330
The white shrimp (Litopenaeus setiferus) is a commercially and recreationally valuable species, yet little is known of its population structure or genetic diversity. White shrimp are distributed along the Atlantic coast of the United States and from the west coast of Florida to the Bay of Campeche, Mexico. In this study, shrimp were collected from North Carolina, South Carolina (four separate collections were taken from 1995 to 1999), Georgia, the Atlantic and Gulf coasts of Florida, Louisiana, Texas and Mexico. DNA was isolated from these individuals, and genetic variation was assessed at six microsatellite loci. These loci were, for the most part, highly polymorphic with an average expected heterozygosity of 0.68. Deviations from Hardy-Weinberg proportions were observed over all samples, but experimental results suggested the presence of null alleles, which confounded a biological interpretation of this result. Pairwise tests of the similarity of allele frequency distributions and distance measure analyses showed broad-scale genetic homogeneity superimposed over occasional indications of random geographical and temporal differentiation. FST and RST estimates over all loci and samples were 0.002 or less and indicated little population structure. Weak but significant genetic differentiation was evident only between pooled western Atlantic and pooled Gulf of Mexico samples. Within the Gulf of Mexico or within the western Atlantic, the large-scale genetic homogeneity observed may be a consequence of genetic mixing resulting from pelagic larvae and adult migrations, while the random local genetic differentiation may be a result of genetic sampling or experimental sampling error. The weak differentiation between shrimp from the Gulf of Mexico and the western Atlantic can be explained by a relatively recent separation of these two populations and/or small amounts of ongoing gene flow.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号