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 31 毫秒
1.
The genome of the social amoeba Dictyostelium discoideum is known to have a very high density of microsatellite repeats, including thousands of triplet microsatellite repeats in coding regions that apparently code for long runs of single amino acids. We used a mutation accumulation study to see if unusually high microsatellite mutation rates contribute to this pattern. There was a modest bias toward mutations that increase repeat number, but because upward mutations were smaller than downward ones, this did not lead to a net average increase in size. Longer microsatellites had higher mutation rates than shorter ones, but did not show greater directional bias. The most striking finding is that the overall mutation rate is the lowest reported for microsatellites: approximately 1 x 10(-6) for 10 dinucleotide loci and 6 x 10(-6) for 52 trinucleotide loci (which were longer). High microsatellite mutation rates therefore do not explain the high incidence of microsatellites. The causal relation may in fact be reversed, with low mutation rates evolving to protect against deleterious fitness effects of mutation at the numerous microsatellites.  相似文献   

2.
Rate and pattern of mutation at microsatellite loci in maize   总被引:30,自引:0,他引:30  
Microsatellites are important tools for plant breeding, genetics, and evolution, but few studies have analyzed their mutation pattern in plants. In this study, we estimated the mutation rate for 142 microsatellite loci in maize (Zea mays subsp. mays) in two different experiments of mutation accumulation. The mutation rate per generation was estimated to be 7.7 x 10(-4) for microsatellites with dinucleotide repeat motifs, with a 95% confidence interval from 5.2 x 10(-4) to 1.1 x 10(-3). For microsatellites with repeat motifs of more than 2 bp in length, no mutations were detected; so we could only estimate the upper 95% confidence limit of 5.1 x 10(-5) for the mutation rate. For dinucleotide repeat microsatellites, we also determined that the variance of change in the number of repeats (sigma(m)2) is 3.2. We sequenced 55 of the 73 observed mutations, and all mutations proved to be changes in the number of repeats in the microsatellite or in mononucleotide tracts flanking the microsatellite. There is a higher probability to mutate to an allele of larger size. There is heterogeneity in the mutation rate among dinucleotide microsatellites and a positive correlation between the number of repeats in the progenitor allele and the mutation rate. The microsatellite-based estimate of the effective population size of maize is more than an order of magnitude less than previously reported values based on nucleotide sequence variation.  相似文献   

3.
Zhang Y  Lu CY  Cao DC  Xu P  Wang S  Li HD  Zhao ZX  Sun XW 《动物学研究》2010,31(5):561-564
利用150个微卫星分子标记在F1代家系的基因型分析过程中,共有27600个等位基因从亲本向子代传递,其中在5个微卫星座位上检测到6个突变的等位基因。对突变的等位基因数目进行统计分析后得出:鲤鱼平均每个世代每个微卫星座位的突变速率为2.53×10-4。在发现突变的5个位点中,经测序发现,突变序列中插入1个以上的重复单元就导致了突变的发生。这些突变表明,鲤鱼的微卫星突变没有遵循严格的渐变突变模型(stepwise mutation model,SMM)。该文关于鲤鱼微卫星突变速率和模式的研究将会对统计鲤鱼有效群体的统计提供有效参数。  相似文献   

4.
Although microsatellites are one of the most popular tools in genetic studies, their mutational dynamics and evolution remain unclear. Here, we apply extensive pedigree genotyping to identify and analyze the patterns and factors associated with de novo germline mutations across nine microsatellite loci in a wild population of lesser kestrels (Falco naumanni). A total of 10 germline mutations events were unambiguously identified in four loci, yielding an average mutation rate of 2.96x10(-3). Across loci, mutation rate was positively correlated with locus variability and average allele size. Mutations were primarily compatible with a stepwise mutation model, although not exclusively involved single-step changes. Unexpectedly, we found an excess of maternally transmitted mutations (male-to-female ratio of 0.1). One of the analyzed loci (Fn2.14) resulted hypermutable (mutation rate=0.87%). This locus showed a size-dependent mutation bias, with longer alleles displaying deletions or additions of a small number of repeat than shorter alleles. Mutation probability at Fn2.14 was higher for females and increased with parental (maternal) age but was not associated with individual physical condition, multilocus heterozygosity, allele length or allele span. Overall, our results do not support the male-biased mutation rate described in other organisms and suggest that mutation dynamics at microsatellite loci are a complex process which requires further research.  相似文献   

5.
We analyze the allelic polymorphisms in seven Y-specific microsatellite loci and a Y-specific alphoid system with 27 variants (alphah I-XXVII), in a total of 89 Y chromosomes carrying the DYS199T allele and belonging to populations representing Amerindian and Na-Dene linguistic groups. Since there are no indications of recurrence for the DYS199C-->T transition, it is assumed that all DYS199T haplotypes derive from a single individual in whom the C-->T mutation occurred for the first time. We identified both the ancestral founder haplotype, 0A, of the DYS199T lineage and seven derived haplogroups diverging from the ancestral one by one to seven mutational steps. The 0A haplotype (5.7% of Native American chromosomes) had the following constitution: DYS199T, alphah II, DYS19/13, DYS389a/10, DYS389b/27, DYS390/24, DYS391/10, DYS392/14, and DYS393/13 (microsatellite alleles are indicated as number of repeats). We analyzed the Y-specific microsatellite mutation rate in 1,743 father-son transmissions, and we pooled our data with data in the literature, to obtain an average mutation rate of.0012. We estimated that the 0A haplotype has an average age of 22,770 years (minimum 13,500 years, maximum 58,700 years). Since the DYS199T allele is found with high frequency in Native American chromosomes, we propose that 0A is one of the most prevalent founder paternal lineages of New World aborigines.  相似文献   

6.
Between three and six tri- and tetranucleotide repeat microsatellite loci were analyzed in 3720 samples collected from four different species of baleen whales. Ten of the 18 species/locus combinations had imperfect allele arrays, i.e., some alleles differed in length by other than simple integer multiples of the basic repeat length. The estimate of the average number of alleles and heterozygosity was higher at loci with imperfect allele arrays relative to those with perfect allele arrays. Nucleotide sequences of 23 different alleles at one tetranucleotide repeat microsatellite locus in fin whales, Balaenoptera physalus, and humpback whales, Megaptera novaeangliae, revealed sequence changes including perfect repeats only, multiple repeats, and partial repeats. The relative rate of the latter two categories of mutation was estimated at 0.024 of the mutation rate involving perfect repeats only. It is hypothesized that single-strand slippage of partial repeats may provide a mechanism for counteracting the continuous expansion of microsatellite loci, which is the logical consequence of recent reports demonstrating directional mutations. Partial-repeat mutations introduce imperfections in the repeat array, which subsequently could reduce the rate of single-strand slippage. Limited computer simulations confirmed this predicted effect of partial-repeat mutations.  相似文献   

7.
Abstract We analyze published data from 592 AC microsatellite loci from 98 species in five vertebrate classes including fish, reptiles, amphibians, birds, and mammals. We use these data to address nine major questions about microsatellite evolution. First, we find that larger genomes do not have more microsatellite loci and therefore reject the hypothesis that microsatellites function primarily to package DNA into chromosomes. Second, we confirm that microsatellite loci are relatively rare in avian genomes, but reject the hypothesis that this is due to physical constraints imposed by flight. Third, we find that microsatellite variation differs among species within classes, possibly relating to population dynamics. Fourth, we reject the hypothesis that microsatellite structure (length, number of alleles, allele dispersion, range in allele sizes) differs between poikilotherms and homeotherms. The difference is found only in fish, which have longer microsatellites and more alleles than the other classes. Fifth, we find that the range in microsatellite allele size at a locus is largely due to the number of alleles and secondarily to allele dispersion. Sixth, length is a major factor influencing mutation rate. Seventh, there is a directional mutation toward an increase in microsatellite length. Eighth, at the species level, microsatellite and allozyme heterozygosity covary and therefore inferences based on large-scale studies of allozyme variation may also reflect microsatellite genetic diversity. Finally, published microsatellite loci (isolated using conventional hybridization methods) provide a biased estimate of the actual mean repeat length of microsatellites in the genome.  相似文献   

8.
Y-chromosome-specific microsatellite variation in Australian aboriginals   总被引:3,自引:0,他引:3  
The frequency distributions of 4 highly polymorphic Y-chromosome-specific microsatellites (DYS19, DYS390, DYS391, and DYS392) were determined in 79 unrelated Australian Aboriginal males from the Northern Territory. These results are compared with those observed in worldwide populations at both the locus and the haplotype level. Common alleles in Aboriginals are DYS19*15 (49%), DYS19*14 (28%), DYS390*19 (39%), DYS390*24 (20%), DYS391*10 (72%), DYS392*11 (63%), and DYS392*13 (28%). No evidence of reduced gene diversity was observed for these Y-chromosome alleles. DYS390 exhibits the most complex arrangement, displaying a bimodal distribution composed of common alleles (*22-*26), and rare short alleles (*18-*20), with an intermediate allele (*21) being absent. DYS390*20, previously reported only in Papuans and Samoans, is observed for the first time in Aboriginals. Compared with a recent study of Aboriginals, our sample exhibits considerable diversity in the haplotypes associated with the rare DYS390*19 allele, indicating that this allele is of considerable antiquity, if it arose as a single deletion event. Combining all 4 Y-chromosome-linked microsatellites produced 41 unique haplotypes, which were linked using a median-joining network. This network shows that most (78%) of our Aboriginal haplotypes fall into 2 distinct clusters, which likely represent 2 separate lineages. Seven haplotypes are shared with haplotypes found in a recent study of Aboriginals, and 7 are shared with a Spanish population. The cluster of Aboriginal haplotypes associated with the short DYS390 alleles does not share any haplotypes with the Spanish, indicating that this cluster of haplotypes is unique to Australian Aboriginals. Limited data from 4 worldwide populations used to construct haplotypes based on 3 loci (DYS19, DYS390, DYS392) show that only 4 of these haplotypes are seen in Australian Aboriginals. Shared haplotypes may be the result of admixture and/or recurrent mutation at these loci. Expanding the haplotype analysis to include biallelic markers on the Y chromosome will resolve this issue.  相似文献   

9.
Uncovering the correct phylogeny of closely related species requires analysis of multiple gene genealogies or, alternatively, genealogies inferred from the multiple alleles found at highly polymorphic loci, such as microsatellites. However, a concern in using microsatellites is that constraints on allele sizes may occur, resulting in homoplasious distributions of alleles, leading to incorrect phylogenies. Seven microsatellites from the pathogenic fungus Coccidioides immitis were sequenced for 20 clinical isolates chosen to represent the known genetic diversity of the pathogen. An organismal phylogeny for C. immitis was inferred from microsatellite-flanking sequence polymorphisms and other restriction fragment length polymorphism-containing loci. Two microsatellite genetic distances were then used to determine phylogenies for C. immitis, and the trees found by these three methods were compared. Congruence between the organismal and microsatellite phylogenies occurred when microsatellite distances were based on simple allele frequency data. However, complex mutation events at some loci made distances based on stepwise mutation models unreliable. Estimates of times of divergence for the two species of C. immitis based on microsatellites were significantly lower than those calculated from flanking sequence, most likely due to constraints on microsatellite allele sizes. Flanking-sequence insertions/deletions significantly decreased the accuracy of genealogical information inferred from microsatellite loci and caused interspecific length homoplasies at one of the seven loci. Our analysis shows that microsatellites are useful phylogenetic markers, although care should be taken to choose loci with appropriate flanking sequences when they are intended for use in evolutionary studies.  相似文献   

10.
Divergent Human Y-Chromosome Microsatellite Evolution Rates   总被引:5,自引:0,他引:5  
In this work, we analyze several characteristics influencing the low variability of the microsatellite DYS19 in the major founder Amerindian Y chromosome lineage containing the point mutation DYS199-T. Variation of DYS19 was compared with that of five other Y-linked tetranucleotide repeat loci (DYS389A, DYS389B, DYS390, DYS391, and DYS393) in the DYS199-T lineage. All the other microsatellites showed significantly higher levels of variability than DYS19 as measured by gene diversity and repeat number variance. Moreover, we had previously shown that DYS19 had high diversity in Brazilians and in several other populations worldwide. Thus, the slow DYS19 evolution in the DYS199-T lineage seems to be both locus and allele specific. To understand the slow DYS19 evolutionary rate, the microsatellite loci were compared according to their mapping on the Y chromosome and also on the basis of structural aspects such as the base composition of the repeat motif and flanking regions and the degree of perfection and size (repeat number) of the variable blocks. The only observed difference that might be related to the low DYS19 variability is its small average number of repeats, a value expected to be closer to the founder DYS19 allele in the DYS199-T lineage. These data were also compared to other derived Y lineages. The Tat-C lineage displayed a lower DYS19 variability correlated to a small average repeat number, while in the DYS234-G lineage, a high DYS19 variability was found associated to a larger average repeat number. This approach reveals that evolution of Y microsatellites in lineages defined by slowly evolving markers, such as point mutations, can be greatly influenced by the size (number of repeats of the variable block) of the founder allele in each microsatellite locus. Thus lineage-dating methods using microsatellite variation should be practiced with great care. Received: 7 November 1998 / Accepted: 9 April 1999  相似文献   

11.
A short tandem repeat-based phylogeny for the human Y chromosome   总被引:9,自引:0,他引:9       下载免费PDF全文
Human Y-chromosomal short tandem repeat (STR) data provide a potential model system for the understanding of autosomal STR mutations in humans and other species. Yet, the reconstruction of STR evolution is rarely attempted, because of the absence of an appropriate methodology. We here develop and validate a phylogenetic-network approach. We have typed 256 Y chromosomes of indigenous descent from Africa, Asia, Europe, Australia, and highland Papua New Guinea, for the STR loci DYS19, DXYS156Y, DYS389, DYS390, DYS392, and DYS393, as well as for five ancient biallelic mutation events: two poly (A) length variants associated with the YAP insertion, two independent SRY-1532 mutations, and the 92R7 mutation. We have used our previously published pedigree data from 11,000 paternity-tested autosomal STR-allele transfers to produce a two-class weighting system for the Y-STR loci that is based on locus lengths and motif lengths. Reduced-median-network analysis yields a phylogeny that is independently supported by the five biallelic mutations, with an error of 6%. We find the earliest branch in our African San (Bushmen) sample. Assuming an age of 20,000 years for the Native American DYS199 T mutation, we estimate a mutation rate of 2.6x10-4 mutations/20 years for slowly mutating Y STRs, approximately 10-fold slower than the published average pedigree rate.  相似文献   

12.
Weetman D  Hauser L  Carvalho GR 《Genetica》2006,127(1-3):285-293
Heterogeneous patterns of microsatellite evolution present a major challenge for the development of mutation models, and an improved understanding of the determinants of variation in mutation rates and patterns among loci, alleles and taxa is required. A 19th Century bottleneck associated with the introduction of clones of the snail Potamopyrgus antipodarum to Britain presented an opportunity to reconstruct recent microsatellite evolution within the most common apomictic lineage. There was significant variation in both the number and step size of mutations among the seven loci studied. Patterns of mutability were consistent with higher mutation rates for di- than trinucleotides and for longer alleles at a locus. Mutation size was influenced in a more complex way, decreasing with relative allele length much more strongly for tri-, than dinucleotides. We found support for this latter, highly novel result in the literature via reanalysis of data in a recent genome-scan study of human microsatellites, which showed a similarly disparate pattern of length-dependence between di- and trinucleotides. In spite of the apomictic form of reproduction and an unusually strong excess of microsatellite contractions in P. antipodarum, there were notable similarities with mutation processes of human microsatellites, supporting the wider taxonomic generality of such evolutionary mechanisms.  相似文献   

13.
The hypermutable nature of some microsatellite loci implies realistic possibilities for the large-scale detection of germline mutations by pedigree analysis. We have developed a model system for mutation analysis by the characterisation of patterns of mutation at three hypervariable microsatellites (two tetranucleotide and one pentanucleotide repeat loci) in barn swallows, all three markers having mutation rates at the percentage level. Here, we study how the mutation rate varies between individual birds of a Spanish population of barn swallows. A total of 53 mutations were identified from 2920 germline transmissions in 90 families with a total of 694 offspring. Mutations were not randomly distributed among individuals (P = 0.020). Attempts to correlate mutation rate with allele size, degree of inbreeding, immunocompetence and male age only revealed a strong effect of allele size. The mean mutation rate differed between colonies of breeding swallows which was probably due to a corresponding variation in allele size between colonies. There was no difference in the mean mutation rate between the Spanish and an Italian population. These results corroborate earlier findings, at the population level, of an allele size effect on the microsatellite mutation rate.  相似文献   

14.
The use of microsatellites in population genetics is hindered by a lack of understanding of the pattern and origin of mutations, the need to develop more specific and better computational models, and a paucity of information about specific taxa and loci. We analyzed between 4 and 10 allele sequences from 10 different microsatellites in Eurasian badgers in order to determine the compliance of the sequences with stepwise mutation models and the origin of that variability which cannot be detected through standard genotyping procedures. All microsatellite loci exhibited imperfections and/or substitutions and indels in the flanking region, as well as additions or deletions of repeat units. Our data set of sequences showed a higher number of imperfect repeats than other published badger and carnivore sequences. This could be attributed to the process of loci isolation because when genetic variability is low, researchers may be more likely to use imperfect loci if these are variable in the population being studied. Locus Mel15 had 2 repetitive arrays: one was part of a polypyrimidine region of a carnivoran short interspersed nuclear element (CAN-SINE) and the other was located in an A-rich region typical of these insertions. In spite of this complexity, heterozygosity was correlated with the maximum number of repeats. Thus, although new theoretical models are being evolved to cover complex patterns of microsatellite mutation, sequencing electromorphs is needed to identify microsatellites or portions of them whose evolution can be modeled under simple models.  相似文献   

15.
A. Estoup  L. Garnery  M. Solignac    J. M. Cornuet 《Genetics》1995,140(2):679-695
Samples from nine populations belonging to three African (intermissa, scutellata and capensis) and four European (mellifera, ligustica, carnica and cecropia) Apis mellifera subspecies were scored for seven microsatellite loci. A large amount of genetic variation (between seven and 30 alleles per locus) was detected. Average heterozygosity and average number of alleles were significantly higher in African than in European subspecies, in agreement with larger effective population sizes in Africa. Microsatellite analyses confirmed that A. mellifera evolved in three distinct and deeply differentiated lineages previously detected by morphological and mitochondrial DNA studies. Dendrogram analysis of workers from a given population indicated that super-sisters cluster together when using a sufficient number of microsatellite data whereas half-sisters do not. An index of classification was derived to summarize the clustering of different taxonomic levels in large phylogenetic trees based on individual genotypes. Finally, individual population X loci data were used to test the adequacy of the two alternative mutation models, the infinite allele model (IAM) and the stepwise mutation models. The better fit overall of the IAM probably results from the majority of the microsatellites used including repeats of two or three different length motifs (compound microsatellites).  相似文献   

16.
Microsatellites are a major component of the human genome, and their evolution has been much studied. However, the evolution of microsatellite flanking sequences has received less attention, with reports of both high and low mutation rates and of a tendency for microsatellites to cluster. From the human genome we generated a database of many thousands of (AC)n flanking sequences within which we searched for common characteristics. Sequences flanking microsatellites of similar length show remarkable levels of convergent evolution, indicating shared mutational biases. These biases extend 25–50 bases either side of the microsatellite and may therefore affect more than 30% of the entire genome. To explore the extent and absolute strength of these effects, we quantified the observed convergence. We also compared homologous human and chimpanzee loci to look for evidence of changes in mutation rate around microsatellites. Most models of DNA sequence evolution assume that mutations are independent and occur randomly. Allowances may be made for sites mutating at different rates and for general mutation biases such as the faster rate of transitions over transversions. Our analysis suggests that these models may be inadequate, in that proximity to even very short microsatellites may alter the rate and distribution of mutations that occur. The elevated local mutation rate combined with sequence convergence, both of which we find evidence for, also provide a possible resolution for the apparently contradictory inferences of mutation rates in microsatellite flanking sequences.  相似文献   

17.
Mutation and evolution of microsatellite loci in Neurospora   总被引:5,自引:0,他引:5  
Dettman JR  Taylor JW 《Genetics》2004,168(3):1231-1248
The patterns of mutation and evolution at 13 microsatellite loci were studied in the filamentous fungal genus Neurospora. First, a detailed investigation was performed on five microsatellite loci by sequencing each microsatellite, together with its nonrepetitive flanking regions, from a set of 147 individuals from eight species of Neurospora. To elucidate the genealogical relationships among microsatellite alleles, repeat number was mapped onto trees constructed from flanking-sequence data. This approach allowed the potentially convergent microsatellite mutations to be placed in the evolutionary context of the less rapidly evolving flanking regions, revealing the complexities of the mutational processes that have generated the allelic diversity conventionally assessed in population genetic studies. In addition to changes in repeat number, frequent substitution mutations within the microsatellites were detected, as were substitutions and insertion/deletions within the flanking regions. By comparing microsatellite and flanking-sequence divergence, clear evidence of interspecific allele length homoplasy and microsatellite mutational saturation was observed, suggesting that these loci are not appropriate for inferring phylogenetic relationships among species. In contrast, little evidence of intraspecific mutational saturation was observed, confirming the utility of these loci for population-level analyses. Frequency distributions of alleles within species were generally consistent with the stepwise mutational model. By comparing variation within species at the microsatellites and the flanking-sequence, estimated microsatellite mutation rates were approximately 2500 times greater than mutation rates of flanking DNA and were consistent with estimates from yeast and fruit flies. A positive relationship between repeat number and variance in repeat number was significant across three genealogical depths, suggesting that longer microsatellite alleles are more mutable than shorter alleles. To test if the observed patterns of microsatellite variation and mutation could be generalized, an additional eight microsatellite loci were characterized and sequenced from a subset of the same Neurospora individuals.  相似文献   

18.
The Colombian jaguar population is thought to contain two different subspecies, Panthera onca centralis and Panthera onca onca. The genetic structure of this population was evaluated using 12 microsatellite loci (n = 62 samples). In addition, 22 jaguar DNA samples from Guatemala, Paraguay, Perú, Bolivia, Venezuela and Brazil were analyzed for these microsatellite loci (n = 84 samples). The results of this study indicate six primary themes. First, the levels of gene diversity were very high. Second, the majority of the loci analyzed showed an absence of Hardy-Weinberg equilibrium, probably due to the Wahlund effect (= population subdivision). Third, several microsatellite loci showed significant heterogeneity between the two supposed subspecies in the country. Nevertheless, gene flow was present between them, and heterogeneity was relatively low, although the assignment analyses showed good classification of the jaguars studied into their respective subspecies. Fourth, the long-term historical effective population sizes were calculated through a maximum likelihood procedure for single and multi-step mutation models. Fifth, seven out of twelve DNA microsatellites studied significantly deviated from a single-step mutation model. However, the overall mean multi-step mutation percentage for these 12 DNA microsatellites was only 6%. Therefore, 94% of mutations were uni-step. Sixth, no bottleneck events were detected in the Colombian jaguar population overall.  相似文献   

19.
Gradients of allele frequencies have long been considered the main genetic characteristic of the European population, but mitochondrial DNA diversity seems to be distributed differently. One Alu insertion (YAP), five tetranucleotide (DYS19, DYS389B, DYS390, DYS391 and DYS393) and one trinucleotide (DYS392) microsatellite loci of the Y chromosome were analysed for geographical patterns in 59 European populations. Spatial correlograms showed clines for most markers, which paralleled the gradients previously observed for two RFLP polymorphisms. Effective separation times between populations were estimated from genetic distances at microsatellite loci. Even after correcting for the possible effects of continuous local gene flow, the most distant Indo-European-speaking populations seem to have separated no more than 7000 years ago. The clinal patterns and the estimated, recent separation times between populations jointly suggest that Y-chromosome diversity in Europe largely reflects a directional demic expansion, which is unlikely to have occurred before the Neolithic period.  相似文献   

20.
Hitchhiking and associative overdominance at a microsatellite locus   总被引:6,自引:2,他引:4  
The possible effects of a selected locus on a closely linked microsatellite locus are discussed and analyzed in terms of coalescent theory and models of the mutation process. Background selection caused by recurrent deleterious mutations will reduce the variance of allele size at a microsatellite locus. The occasional substitution of advantageous alleles (genetic hitchhiking) will also reduce the variance, but a high mutation rate at a microsatellite locus can restore the variance relatively rapidly. Overdominance at the selected locus will increase the variance at the microsatellite locus and create partitioning of the variation in allele size among gametes carrying one or the other of the overdominant alleles. These results suggest that neutral microsatellite loci can provide indicators of selective processes at closely linked loci.   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