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 15 毫秒
1.
Epperson BK  Allard RW 《Genetics》1987,115(2):341-352
Pairwise linkage disequilibrium values (D) were estimated for 14 allozyme loci in two natural populations of lodgepole pine (Pinus contorta ssp. latifolia). Maternal multilocus genotypes were inferred from samples of (haploid) megagametophytic seed-endosperms. Coupling/repulsion double heterozygotes were distinguished for closely linked pairs of loci. Assays of seven of the loci in seed embryos allowed estimates of D for these loci in the outcross pollen pool (estimates of outcrossing rates indicate no significant departures from random mating in either population). No disequilibrium was observed between unlinked loci in either maternal genotypes or outcross pollen. However, significant disequilibrium was observed within and between gametes for some allelic combinations of four tightly linked loci; the assumption of random association of gamete types within individuals is thus invalid for some loci in lodgepole pine. Possible causes of the observed D were examined using the noncentrality parameter of the general noncentral chi square distribution. We concluded, from estimates of population size, linkage and measurements of population substructure, that neither drift nor population subdivision was responsible for the significant values of D which were observed and that epistatic selection was the most likely cause of the disequilibrium observed.  相似文献   

2.
The effect of genotype, the origin of genotype, and germination temperature on Scots pine pollen grain size, hydration rate, germinability, and tube growth was studied in vitro. The mean sizes of dry and germinated pollen grains varied among pollen genotypes in different ways, thus the hydration rate varied among genotypes. Pollen from Scots pine that originates in northern Finland hydrated more than pollen from a population in southern Finland. Germination temperature had no effect on the hydration rate. Germinability and tube growth rate of northern genotypes were higher at 20 °C than at 15 °C. Differences among southern genotypes were not significant. At 15 °C, the germinability and pollen tube growth rate of northern genotypes were lower than southern genotypes. At 20 °C, the differences were not significant. It appears that germination and growth of pollen from northern populations are enhanced at higher temperatures whereas pollen from southern populations is unaffected.  相似文献   

3.
Spatial autocorrelation analyses of point samples within two populations of lodgepole pine (Pinus contorta ssp. latifolia) indicate that single-locus mature tree and pollen genotypes are distributed in a nearly random fashion for most of the allozyme loci assayed. This lack of structure in the distributions of most genotypes is consistent with outcrossing rates that are very nearly 1.0 and with estimates indicating that both pollen and seed are dispersed over long distances in lodgepole pine. However, spatial autocorrelation of genotypes for a few loci suggests that genotypes at these loci may be under natural selection.  相似文献   

4.
The Canary Island pine weevil Brachyderes rugatus (Wollaston) consists of four allopatric subspecies that are thought to have arisen from several historic colonization events within the archipelago. We have isolated and optimized seven microsatellite loci from Brachyderes rugatus calvus from Gran Canaria. Six of these loci are polymorphic within B. rugatus (11–22 alleles per locus; heterozygosity between 0.43 and 0.84). There is no evidence for heterozygote deficit within populations or for linkage disequilibrium between pairs of loci. These molecular markers are likely to prove useful tools for quantifying the genetic variability of bottlenecked island populations.  相似文献   

5.
We evaluated the population structure of the bog and dry land populations of the Siberian pine Pinus sibirica (P. sibrica) in Western Siberia using nuclear genome markers. Six pairs of nuclear microsatellite loci were used for this analysis. We detected 30 allelic variants in 120 individuals of four populations of P. sibirica. We established that the studied populations differ by genetic structure. The most essential differences were identified between the Siberian pine population from oligotrophic bog and the group of populations from dry land within eutrophic bogs and near settlements P. sibirica forest (F ST = 0.019; D N = 0.053). We estimated that diversification of the West Siberian populations of P. sibirica exceeded 2.4% (F ST = 0.024), based on an analysis of SSR markers.  相似文献   

6.
On range-wide and regional scales, climate and site factors exert control over tree growth, masking the genetic basis of biomass accumulation and allocation. To determine intrinsic population differences in productivity, aboveground net primary production (ANPP) was measured in 16-year-old Scots pine from 19 geographically distinct populations grown in a common garden experiment in central Poland (52°N). The populations originated from the northern (>55°N), central (54–47°N), and southern (<45°N) European range of Scots pine. We calculated ANPP from aboveground growth components, using diameter-based allometric equations developed for this site. Average foliage, aboveground woody and total ANPP differed significantly among populations and were greater for central European populations than for the southern and northern ones. Stocking and total ANPP per tree were positively correlated to stand aboveground biomass (r 2≥0.71). The relationship between the latitude of seed origin and ANPP was curvilinear and maximum for populations originating near the planting site (52°N). ANPP declined in populations with increasing longitude eastward from the Atlantic Ocean towards the center of the continent. This study underscores the potentially large genetic control of ANPP and biomass accumulation among diverse Scots pine populations. Received: 3 January 2000 / Accepted: 29 March 2000  相似文献   

7.
In the context of a general survey on genetic variation of isozyme-gene systems which function in the carbohydrate degradation and conversion, we detected a reciprocal relationship between genetic diversity at the hexokinase (HEK-A) and phosphoglucose isomerase (PGI-B) loci in Scots pine populations. Further studies on Norway spruce, Douglas-fir and Siberian stone pine revealed that this relationship appears to be a more general phenomenon in conifers such that increasing diversity at one locus is correlated with a decrease in diversity at the other locus. Since the two gene loci are not structurally linked but are encoding enzymes of two sucessive metabolic steps in the glucose conversion towards glycolysis, it is assumed that some sort of selection, especially during germination and early embryo development, may be the causal explanation. A metabolically-based model incorporating selective advantage and disadvantage of alternate two-locus genotypes at HEK-A/PGI-B was presented in order to elucidate the possible adaptive nature of this reciprocal relationship. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

8.
In the present study we investigated the genetic structure and genetic diversity of Pinus sylvestris populations in Bulgaria using chloroplast microsatellite markers and terpene analysis. We were interested in addressing the following questions: (1) can population structure in Scots pine be detected via chloroplast microsatellites markers and terpenes; (2) are there differences in population differentiation between the two analyses; and (3) how are the patterns related to geographic distances. Twelve provenances were chosen throughout the species' range in Bulgaria. Following DNA extraction, chloroplast microsatellite (cpSSR) loci were surveyed using six primer pairs. Between 4 to 8 size variants were identified at each locus. A total of 35 size variants at the six loci were identified, 11 occurring at low frequencies (<1%). They were combined in 134 different haplotypes, of which seven represent 1/3 of the genetic structure. AMOVA analysis revealed that 10.99% of the variation was found among populations, while 89.01% was expressed within populations. The cpSSR analysis divided Scots pine populations into two groups, the first represented by populations located in the south-western part of the Rhodopes and Pirin mountains, while the second group is located in the northeast of Rhodopes and Rila mountains. Terpene analysis revealed that on average, 53% of the monoterpene pool in P. sylvestris was accounted for by -pinene (range 47–59%) followed by β-pinene (range 6–12%). The presence of two distinct groups is weekly consistent with physical distances between populations, similar significant correlation between genetic distance determined by chloroplast microsatellites analysis and chemotype distance (determined by terpenes) was observed. Our results suggest that the structural pattern of genetic diversity of cpDNA in Scots pine populations is the consequence of historical biogeographic processes.  相似文献   

9.
In the present study we investigated the genetic structure and genetic diversity of Pinus heldreichii populations in Bulgaria using chloroplast microsatellite markers and terpene analysis. We were interested in addressing the following questions: (1) can population structuring in Bosnian pine be detected via chloroplast microsatellite markers; (2) are there differences in population differentiation as determined by terpenes and microsatellites; and (3) how are the patterns of size variant frequencies and geographical distances related. Four provenances were chosen throughout the species' range in Bulgaria. Following DNA extraction, chloroplast microsatellite (cpSSR) loci were surveyed using 6 primer pairs. Between 2 and 5 size variants were identified at each locus. A total of 16 size variants at the 6 loci were identified, 4 occurring at low frequencies. They were combined in 21 different haplotypes including 11 that were unique. AMOVA analysis revealed that 18.25% of the variation was found among populations, while 81.75% was expressed within populations. The cpSSR analysis divided Bosnian pine populations into two groups, the first represented by populations B, C and D located in the south and north-western part of the Pirin and Slavianka mountains, while the second group, represented by population A, is located in the north-eastern Pirin mountain. Terpene analysis revealed that on average, 59% of the monoterpene pool in P. heldreichii is accounted for by limonene (range 36–48%) followed by α-pinene (range 16–17%). The presence of two distinct groups (Pop-A, Pop-D and Pop-B, Pop-C) is more consistent with physical distances between populations. No significant correlation between genetic distance determined by chloroplast microsatellites analysis and chemotype distance determined by terpenes was observed. Our results suggest that the structural pattern of genetic diversity of cpDNA in Bosnian pine populations is the consequence of historical and biogeographical processes.  相似文献   

10.
Random amplified polymorphic DNAs (RAPDs) were used to construct linkage maps of the parent of a longleaf pine (Pinus palustris Mill.) slash pine (Pinus elliottii Englm.) F1 family. A total of 247 segregating loci [233 (1∶1), 14 (3∶1)] and 87 polymorphic (between parents), but non-segregating, loci were identified. The 233 loci segregating 1∶1 (testcross configuration) were used to construct parent-specific linkage maps, 132 for the longleaf-pine parent and 101 for the slash-pine parent. The resulting linkage maps consisted of 122 marker loci in 18 groups (three or more loci) and three pairs (1367.5 cM) for longleaf pine, and 91 marker loci in 13 groups and six pairs for slash pine (952.9 cM). Genome size estimates based on two-point linkage data ranged from 2348 to 2392 cM for longleaf pine, and from 2292 to 2372 cM for slash pine. Linkage of 3∶1 loci to testcross loci in each of the parental maps was used to infer further linkages within maps, as well as potentially homologous counterparts between maps. Three of the longleaf-pine linkage groups appear to be potentially homologous counterparts to four different slash-pine linkage groups. The number of heterozygous loci (previously testcross in parents) per F1 individual, ranged from 96 to 130. With the 87 polymorphic, but non-segregating, loci that should also be heterozygous in the F1 progeny, a maximum of 183–217 heterozygous loci could be available for mapping early height growth (EHG) loci and for applying genomic selection in backcross populations.  相似文献   

11.
Seed longevity varies considerably in cultivated rice, but the underlying mechanism of longevity is not well understood. To measure seed longevity, we performed an aging treatment at 45 °C on seeds maintained at 14 % moisture content for 14 days. We measured the percentage germination of both treated and normal seeds at 25 °C as a control of seed longevity using four replications over 2 years. In total, 140 accessions from a core collection with diverse origins were genotyped using 204 SSR markers, which distributed into 12 chromosomes, to identify marker–trait associations with seed longevity. An analysis of the population structure revealed four subgroups. The r 2 values ranged from 0.0 to 0.8901 for all intrachromosomal loci pairs, with an average of 0.0773. Linkage disequilibrium (LD) between linked markers decreased with distance and displayed a substantial drop in LD decay values between 20 and 50 cM. Marker–trait associations were investigated using a mixed linear model approach, considering both population structure (Q) and kinship (K). Twelve marker–trait associations (P < 0.01) were common between the two germination treatments and over the 2-year study, explaining more than 10 % of the total variation. These ten different markers were distributed on five chromosomes. The significant associated SSR markers identified will be useful to seed-bank managers to ensure collections are maintained at high levels of viability to avoid loss of genotypes from the population and for marker-assisted selection.  相似文献   

12.
In bivalves, heterozygote deficiencies and departures from Hardy-Weinberg equilibrium (HWE) in microsatellite analysis are common and mainly attributed to inbreeding, genetic patchiness (Walhund effect), or null alleles. We checked for the occurrence of null alleles at 3 microsatellite loci in 3 populations of black-lipped pearl oyster, Pinctada margaritifera, using a step-by-step method to re-amplify homozygotes and null individuals with redesigned primer pair combinations. After amplification with original primer pairs, the 3 populations exhibited null alleles, absence of structure, and significant departure from HWE for all 3 loci due to heterozygote deficiencies. After 3 re-amplification steps, with modified primer sets, all loci were corrected for null alleles. Once corrected, all populations appeared at HWE, demonstrating that null alleles were responsible for the initial disequilibrium of the populations. Furthermore, analysis from corrected genotypes demonstrates significant genetic differentiation for one population from the other 2.  相似文献   

13.
In order to study genetic diversity of white birch (Betula platyphylla), 544 primer pairs were designed based on the genome-wide Solexa sequences. Among them, 215 primer pairs showed polymorphism between five genotypes and 111 primer pairs that presented clear visible bands in genotyping 41 white birch plants that were collected from 6 different geographical regions. A total of 717 alleles were obtained at 111 loci with a range of 2 to 12 alleles per locus. The results of statistic analysis showed that polymorphic frequency of the alleles ranged from 17% to 100% with a mean of 55.85%; polymorphism information content (PIC) of the loci was from 0.09 to 0.58 with a mean of 0.30; and gene diversity between the tested genotypes was from 0.01 to 0.66 with a mean of 0.36. The results also indicated that major allele frequency ranged from 0.39 to 1.00 with an mean of 0.75; expected heterozygosity from 0.22 to 0.54 with a mean of 0.46; observed heterozygosity from 0.02 to 0.95 with a mean of 0.26; Nei''s index from 0.21 to 0.54 with a mean of 0.46; and Shannon''s Information from 0.26 to 0.87 with a mean of 0.66. The 41 white birch genotypes at the 111 selected SSR loci showed low to moderate similarity (0.025-0.610), indicating complicated genetic diversity among the white birch collections. The UPGMA-based clustering analysis of the allelic constitution of 41 white birch genotypes at 111 SSR loci suggested that the six different geographical regions can be further separated into four clusters at a similarity coefficient of 0.22. Genotypes from Huanren and Liangshui provenances were grouped into Cluster I, genotypes from Xiaobeihu and Qingyuan provenances into Cluster II, genotypes from Finland provenance into Cluster III, and genotypes from Maoershan into Cluster IV. The information provided in this study could help for genetic improvement and germplasm conservation, evaluation and utilization in white birch tree breeding program.  相似文献   

14.
We developed eight polymorphic nuclear microsatellite markers for the Swiss stone pine (Pinus cembra L.), of which seven may be amplified in a multiplex polymerase chain reaction. Allelic polymorphism across all loci and 40 individuals representing two populations in the Swiss Alps was high (mean = 7.6 alleles). No significant linkage disequlibrium was displayed between pairs of loci. Significant deviation from Hardy–Weinberg equilibrium was revealed at three loci in one population. Cross–amplification was achieved in two related species within the genus (P. sibirica and P. pumila). Thus, the markers may be useful for population genetic studies in these three pine species. They will be applied in ongoing projects on genetic diversity and patterns of gene flow in P. cembra.  相似文献   

15.
16.
Linkage analysis is commonly used to find marker-trait associations within the full-sib families of forest tree and other species. Study of marker-trait associations at the population level is termed linkage-disequilibrium (LD) mapping. A female-tester design comprising 200 full-sib families generated by crossing 40 pollen parents with five female parents was used to assess the relationship between the marker-allele frequency classes obtained from parental genotypes at SSR marker loci and the full-sib family performance (average predicted breeding value of two parents) in radiata pine (Pinus radiata D. Don). For alleles (at a marker locus) that showed significant association, the copy number of that allele in the parents was significantly correlated, either positively or negatively, with the full-sib family performance for various economic traits. Regression of parental breeding value on its genotype at marker loci revealed that most of the markers that showed significant association with full-sib family performance were not significantly associated with the parental breeding values. This suggests that over-representation of the female parents in our sample of 200 full-sib families could have biased the process of detecting marker-trait associations. The evidence for the existence of marker-trait LD in the population studied is rather weak and would require further testing. The exact test for genotypic disequilibrium between pairs of linked or unlinked marker loci revealed non-significant LD. Observed genotypic frequencies at several marker loci were significantly different from the expected Hardy-Weinberg equilibrium. The possibilities of utilising marker-trait associations for early selection, among-family selection and selecting parents for the next generation of breeding are also discussed.  相似文献   

17.
We isolated and characterized eight novel microsatellite loci in the southern emu‐wren (Stipiturus malachurus). We used nonradioactive polymerase chain reaction (PCR)‐based techniques to screen an enriched genomic DNA library. Based on genotypes from a single population, six loci showed no evidence of null alleles and were polymorphic (allele range = 2–9, mean heterozygosity = 0.57), and one locus was sex‐linked (NA = 4). These loci were variable and had different allele size ranges in three other populations of southern emu‐wrens, and are therefore useful for determining levels of genetic diversity within and between populations of the species.  相似文献   

18.
E. Zouros  C. B. Krimbas  S. Tsakas    M. Loukas 《Genetics》1974,78(4):1223-1244
Gametic frequencies in one mainland and one island population of D. subobscura were obtained by means of extracting wild chromosomes and subsequently analyzing them for inversions and allozymes. The high degree of cytological heterogeneity which characterizes these populations is not reflected in the genetic data. Two cases of non-random association were observed among eighteen pair-wise comparisons involving gene alleles and inversions to which the locus is linked. In both cases exchange of alleles at the locus is completely suppressed by the inversions. Four cases of linkage disequilibrium were detected among eighteen pairs of loci; two of them could best be explained as transient associations generated by random drift. The results suggest that disequilibria among enzyme loci are not widespread in natural populations—Populations with a lower degree of chromosomal variation are genetically as variable as populations with a higher degree of chromosomal variation. This observation does not support the hypothesis that selection in marginal homokaryotypic populations is for specialized homozygous genotypes.  相似文献   

19.
Abstract 1 The European pine sawfly, Neodiprion sertifer (Geoffroy) (Hymenoptera, Diprionidae), frequently defoliates Scots pine (Pinus sylvestris L.) forests in northern Europe. It overwinters as an egg. It has been proposed that the high egg mortality caused by low winter temperatures limits the occurrence of outbreaks to the southern part of Fennoscandia. 2 In this study, variation in freezing avoidance by egg supercooling between four Finnish populations (originating between latitudes 60°N and 69°N) of N. sertifer was tested by differential thermal analysis. Offspring of 20 females within each population were selected for the study. The freezing avoidance of parasitized eggs was also examined. 3 The northernmost Inari population was found to be the cold hardiest, and the southernmost (Hanko) was the least hardy population. The within‐population variation between females was greatest in the population from Inari, and the next greatest in the one from Hanko. The inland populations in Eastern Finland had the smallest within‐population variation in freezing avoidance. 4 The high variation in freezing avoidance of eggs will enable N. sertifer to adapt to the predicted climate change and to spread its distribution northwards. This may also change the risk for outbreaks in this area. Parasitized eggs froze at higher temperature than healthy eggs. This observation indicates that N. sertifer may experience reduced egg parasitism in certain winter climate conditions.  相似文献   

20.
We have reconstructed partial genealogies in a sample of 44 SW Amazonian Rondonian Surui, in which 45 dinucleotide short tandem repeat polymorphisms had previously been typed. The genotypes of 488 pairs of individuals having an age difference of 13 or greater were compared, and parentage was excluded if a pair failed to share an allele at more than one locus. In order to test the power of this method, we computed the expected distribution of the number of exclusionary loci for such pairs of unrelated individuals, as well as that for individuals with different degrees of relatedness, and compared it to the observed distribution. We estimated that the pairs compared contained ∼20% of individual pairs with a first-cousin relation or closer. A total of 25 pairs were identified as possible parent-child. In three instances, we could identify two or more children having a common parent; we computed a relatedness coefficient in order to establish whether the children were full or half sibs. The genealogies inferred show that instances of polygyny and polyandry (or, alternatively, serial mating), in addition to apparent monogamy, can be found among the Surui. The Surui sample can be used as a model for paleoanthropological populations, in which the determination of relatedness can provide further insights into the social structure of past populations. We estimate that, depending on the history of the populations and the degree of inbreeding, 10–20 highly informative nuclear loci should be typed in order to infer genealogies with acceptable confidence. Am J Phys Anthropol 108:137–146, 1999. © 1998 Wiley-Liss, Inc.  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