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.
Hughes AL 《Immunogenetics》2000,51(6):473-486
 The phylogenetic relationships and patterns of nucleotide substitution were compared for introns and exons of class II major histocompatibility complex (MHC) genes in three datasets: human DRB1, human DQA1, and cyprinid fish DAB1. In both human DRB1 and cyprinid DAB1, there was strong evidence that recombination events between alleles have occurred in such a way that intron and exon sequences of a given allele do not necessarily share the same evolutionary history. In the case of human DRB1, recombination was found to have homogenized intron 1 and intron 2 sequences relative to exon 2 sequences within lineages of alleles but not between lineages. As a result, mean divergence times of intron sequences are much more recent than those of exonic sequences. Thus, the divergence time of DRB1 introns cannot be used to date that of exons in the same alleles, and the hypothesis that most human DRB1 polymorphism is of very recent origin is not supported. Received: 5 September 1999 / Revised: 30 December 1999  相似文献   

2.
Background and aim: Polymorphisms of Helicobacter pylori cagA and vacA genes do exist and may contribute to differences in H. pylori infection and gastroduodenal diseases among races in the Malaysian population. This study was conducted to characterize the polymorphisms in H. pylori cagA and vacA in Malaysian population. Methods: A total of 110 H. pylori isolates were genotyped by PCR and sequenced for cagA and PCR‐RFLP for vacA. Results: East Asian cagA was predominantly detected (64.5%), whereas vacA s1m1 and s1m2 alleles were detected in 60.9 and 37.3% of strains, respectively. A statistical association between cagA type with patients’ ethnicity (p < .0001) and age group >50 years old (p = .027) was identified. vacA alleles showed significant association with age group >50 years old (p = .017) and increased neutrophil activity in gastric mucosa (p = .028 and p = .016 for moderate and marked activity, respectively). Further identification of vacA polymorphism revealed that 84% of strains from Malays and Indians showed one RFLP pattern (RFLP‐1), whereas more than one RFLP patterns (RFLP‐2, 3, 4, 5, 6, and 8) were predominantly observed in strains from Chinese (82%) (p < .0001). Increasing severity of gastric inflammation was observed in gastric mucosa infected with strains carrying RFLP‐2, 3, 4, 5, and 6 (p = .037). About 86.6% of H. pylori strains with East Asian cagA were vacA RFLP‐2, 3, 4, 5, 6, and 8, and 88% of Western cagA strains were vacA RFLP‐1 (p < .0001). Chinese and Indians are susceptible to different virulence genotypes of H. pylori, whereas Malays showed a mixed virulence genotypes. Conclusion: Marked differences in the polymorphisms of cagA and vacA were observed among strains in Malaysian population. This provides a new insight into the pathogenicity of H. pylori in multiracial population.  相似文献   

3.
Data on the geographic prevalence of Helicobacter pylori iceA and babA alleles in Eastern Europe are still relatively scant. The aim of this study was to evaluate the prevalence of iceA and babA genotypes in Bulgarian symptomatic patients. The iceA and babA genotypes were evaluated by PCR with pure cultures in strains from 196 and 181 patients, respectively. Mixed infections were found in 10.2% of all 196 patients. Prevalence of H. pylori genotypes in patients with single-strain infections was 69.3% for iceA1, 30.7% for iceA2, 82.4% for cagA +, 89.2% for vacA s1, 10.8% for vacA s2, 39.8% for vacA m1, 60.2% for vacA m2 and 48.8% for babA2. Within the iceA1 positive strains, 94.3% and 88.5% were also vacA s1a and cagA positive, respectively. Of the babA2 positive strains, 100.0%, 92.4% and 72.2% were also vacA s1a, cagA and iceA1 positive, respectively. Ulcer patients had more often strains with cagA positive status and vacA s1a allele. Although neither iceA1 nor babA2 were more common in ulcer patients, the combination of both alleles was more frequent (48.1%) in the ulcer patients than in the rest (28.7%). Clarithromycin susceptible strains had more often iceA1 allele (74.4%) than the resistant strains (55.3%). In conclusion, the results demonstrated a high prevalence of virulent H. pylori in Bulgaria. Both iceA1 and babA2 genotypes were associated with other virulence factors of H. pylori and, in addition, the iceA1 allele was associated with the strain susceptibility.  相似文献   

4.
Helicobacter pylori is a gram-negative bacterium that colonizes the stomach of nearly half of the world''s population. Genotypic characterization of H. pylori strains involves the analysis of virulence-associated genes, such as vacA, which has multiple alleles. Previous phylogenetic analyses have revealed a connection between modern H. pylori strains and the movement of ancient human populations. In this study, H. pylori DNA was amplified from the stomach tissue of the Kwäday Dän Ts''ìnchi individual. This ancient individual was recovered from the Samuel Glacier in Tatshenshini-Alsek Park, British Columbia, Canada on the traditional territory of the Champagne and Aishihik First Nations and radiocarbon dated to a timeframe of approximately AD 1670 to 1850. This is the first ancient H. pylori strain to be characterized with vacA sequence data. The Tatshenshini H. pylori strain has a potential hybrid vacA m2a/m1d middle (m) region allele and a vacA s2 signal (s) region allele. A vacA s2 allele is more commonly identified with Western strains, and this suggests that European strains were present in northwestern Canada during the ancient individual''s time. Phylogenetic analysis indicated that the vacA m1d region of the ancient strain clusters with previously published novel Native American strains that are closely related to Asian strains. This indicates a past connection between the Kwäday Dän Ts''ìnchi individual and the ancestors who arrived in the New World thousands of years ago.  相似文献   

5.
Microsatellite length variation was investigated at a highly variable microsatellite locus in four species of Apodemus. Information obtained from microsatellite allele sequences was contrasted with allele sizes, which included 18 electromorphs. Additional analysis of a 400-bp unique sequence in the flanking region identified 26 different haplotype sequences or ``true' alleles in the sample. Three molecular mechanisms, namely, (1) addition/deletion of repeats, (2) substitutions and indels in the flanking region, and (3) mutations interrupting the repeat, contributed to the generation of allelic variation. Size homoplasy can be inferred for alleles within populations, from different populations of the same species, and from different species. We propose that microsatellite flanking sequences may be informative markers for investigating mutation processes in microsatellite repeats as well as phylogenetic relationships among alleles, populations, and species. Received: 3 November 1999 / Accepted: 2 May 2000  相似文献   

6.
The human gastric pathogen, Helicobacter pylori, has co-evolved with its host and established itself in the human stomach possibly millions of years ago. Therefore, the diversity of this bacterium is important in its clinical manifestations. Our aim has been to evaluate the genetic diversity of 40 H. pylori clinical isolates from four different parts of China. The methods of multi-locus sequence typing and vacA allele genotyping were used to assess their genetic diversity. To discriminate MLST, the vacA genotype method was used to identify strains. Patients from the northern, eastern, southern, and southwestern parts of China were recruited randomly from the cities of Beijing, Shanghai, Guangzhou, and Chongqing, respectively. Most of the sequence types are new and have never been reported in the database of the H. pylori multi-locus sequence typing system. The most prevalent vacA genotype in patients was s1a/m2 (80.0%), followed by s1b/m2 (17.5%). In contrast, the s1a/m1 genotype was scarcely represented (2.5%). The vacA genotype varied for each ST. These results showed that the MLST method offers high resolution of the H. pylori isolates in China when compared to vacA genotyping. The vacA allelic s1a has been correlated with the peptic ulcer. Because of the paucity of data on human isolates due to the absence of systematic investigations of H. pylori in China, the data provide useful information for understanding the epidemiology of H. pylori in China from the viewpoint of nucleotide sequence databases.  相似文献   

7.
Background: Helicobacter pylori infection is usually acquired in childhood, but little is known about its natural history in asymptomatic children, primarily due to the paucity of non‐invasive diagnostic methods. H. pylori strains harboring cagA and specific alleles of hopQ and vacA are associated with increased risk for gastric cancer. Many studies of H. pylori virulence markers in children have the bias that symptomatic subjects are selected for endoscopy, and these children may harbor the most virulent strains. Our aim is to genotype cagA, hopQ, and vacA alleles in stool DNA samples of healthy Colombian children residing in an area with high incidence of gastric cancer, to avoid selection bias resulting from endoscopy. Methods: H. pylori status of 86 asymptomatic children was assessed by 13C‐urea breath test (UBT) and PCR. H. pylori 16S rRNA, cagA, hopQ, and vacA genes were amplified from stool DNA samples and sequenced. Results: UBT was positive in 69 (80.2%) of 86 children; in stool DNA analysis, 78.3% were positive by 16S rRNA PCR. cagA, vacA, and hopQ were detected in 66.1%, 84.6%, and 72.3% of stool DNA samples from 16S rRNA‐positive children. Of the children’s DNA samples, which revealed vacA and hopQ alleles, 91.7% showed vacA s1 and 73.7% showed type I hopQ. Type I hopQ alleles were associated with cagA positivity and vacA s1 genotypes (p < 0.0001). Conclusions: Using stool DNA samples, virulence markers of H. pylori were successfully genotyped in a high percentage of the asymptomatic infected children, revealing a high prevalence of genotypes associated with virulence. Type I hopQ alleles were associated with the presence of cagA and the vacA s1 genotype.  相似文献   

8.
We previously reported the occurrence of genetically‐diverse symbiotic dinoflagellates (zooxanthellae) within and between 7 giant clam species (Tridacnidae) from the Philippines based on the algal isolates' allozyme and random amplified polymorphic DNA (RAPD) patterns. We also reported that these isolates all belong to clade A of the Symbiodinium phylogeny with identical 18S rDNA sequences. Here we extend the genetic characterization of Symbiodinium isolates from giant clams and propose that they are conspecific. We used the combined DNA sequences of the internal transcribed spacer (ITS)1, 5.8S rDNA, and ITS2 regions (rDNA‐ITS region) because the ITS1 and ITS2 regions evolve faster than 18S rDNA and have been shown to be useful in distinguishing strains of other dinoflagellates. DGGE of the most variable segment of the rDNA‐ITS region, ITS1, from clonal representatives of clades A, B, and C showed minimal intragenomic variation. The rDNA‐ITS region shows similar phylogenetic relationships between Symbiodinium isolates from symbiotic bivalves and some cnidarians as does 18S rDNA, and that there are not many different clade A species or strains among cultured zooxanthellae (CZ) from giant clams. The CZ from giant clams had virtually identical sequences, with only a single nucleotide difference in the ITS2 region separating two groups of isolates. These data suggest that there is one CZ species and perhaps two CZ strains, each CZ strain containing individuals that have diverse allozyme and RAPD genotypes. The CZ isolated from giant clams from different areas in the Philippines (21 isolates, 7 clam species), the Australian Great Barrier Reef (1 isolate, 1 clam species), Palau (8 isolates, 7 clam species), and Okinawa, Japan (1 isolate, 1 clam species) shared the same rDNA‐ITS sequences. Furthermore, analysis of fresh isolates from giant clams collected from these geographical areas shows that these bivalves also host indistinguishable clade C symbionts. These data demonstrate that conspecific Symbiodinium genotypes, particularly clade A symbionts, are distributed in giant clams throughout the Indo‐Pacific.  相似文献   

9.
The mutational mechanism underlying the striking diversity in MHC (major histocompatibility complex) genes in vertebrates is still controversial. In order to evaluate the role of inter- and intragenic recombination in MHC gene diversification, we examined patterns of nucleotide polymorphism across an exon/intron boundary in a sample of 31 MHC class IIB sequences of three-spined stickleback (Gasterosteus aculeatus). MHC class IIB genes of G. aculeatus were previously shown to be under diversifying (positive) selection in mate choice and pathogen selection experiments. Based on recoding of alignment gaps, complete intron 2 sequences were grouped into three clusters using maximum-parsimony analysis. Two of these groups had >90% bootstrap support and were tentatively assigned single locus status. Intron nucleotide diversity within and among loci was low (p-distance within and among groups = 0.016 and 0.019, respectively) and fourfold lower than the rate of silent mutations in exon 2, suggesting that noncoding regions are homogenized by frequent interlocus recombination. A substitution analysis using GENECONV revealed as many intergenic conversion events as intragenic ones. Recombination between loci may explain the occurrence of sequence variants that are particularly divergent, as is the case in three-spined stickleback, with nucleotide diversity attaining dN = 0.39 (peptide-binding residues only). For both MHC class II loci we also estimated the amount of intragenic recombination as population rate (4Ner) under the coalescent and found it to be approximately three times higher compared to point mutations (Watterson estimate per gene, 4Neμ). Nonindependence of molecular evolution across loci and frequent recombination suggest that MHC class II genes of bony fish may follow different evolutionary dynamics than those of mammals. Our finding of widespread recombination suggests that phylogenies of MHC genes should not be based on coding segments but rather on noncoding introns. [Reviewing Editor: Dr. Richard Kliman]  相似文献   

10.
Morphological, toxicological, and genetic variation was examined among 19 strains of Nodularia. The strains examined could be morphologically discriminated into four groups corresponding to N. spumigena Mertens, N. sphaerocarpa Bornet et Flahault, and two strains that did not clearly correspond to currently accepted Nodularia species. Genetic variation was examined using nucleotide sequencing of the phycocyanin intergenic spacer region (cpcBA-IGS) and RAPD-PCR. The PCR-RFLP of the cpcBA-IGS differentiated four genotypes corresponding to the four morphological groups. However, nucleotide sequencing of 598 bp of the 690-bp fragment showed that one of the three strains corresponding to N. sphaerocarpa (PCC 7804) was genetically divergent from the other two, suggesting that it constitutes a distinct species. Nucleotide variation within the morphospecies groups was limited (<1%), and all 14 Australian strains of N. spumigena possessed identical cpcBA-IGS sequences. The RAPD-PCR differentiated the same groups as the cpcBA sequencing and discriminated each of the seven different Australian populations of N. spumigena. Strains from within a bloom appeared genetically identical; however, strains isolated from different blooms could be separated into either a western or a southeastern Australian cluster, with one strain from western Australia showing considerable genetic divergence. The pattern of variation suggests that individual blooms of N. spumigena are clonal but also that Australian N. spumigena populations are genetically distinct from each other. Examination of genetic distance within and between blooms and within and between morphological groups showed clear genetic dicontinuities that, in combination with the cpcBA-IGS data, suggest that Nodularia contains genetically distinct morphospecies rather than a continuous cline of genetic variation. Furthermore, these morphospecies are genetically variable, exhibiting hierarchical patterns of genetic variation on regional and global scales. Production of the hepatotoxin nodularin was not restricted to one genetic lineage but was distributed across three of the five genotypic groups. A strain of N. spumigena from a nontoxic Australian population was found to fall within the range of genetic variation for other toxic Australian strains and appears to be a unique nontoxic strain that might have arisen by loss of toxin production capacity.  相似文献   

11.
The 3′‐terminal sequences (c. 1700 nt) of the RNA genome of 10 Turnip mosaic virus (TuMV) isolates from different hosts in Zhejiang province, China, were determined. Phylogenetic analysis of the coat protein nucleotide sequences revealed that most TuMV sequences fell into two distinct clusters. The Chinese isolates B1‐B4 (from Brassica spp.) were similar and placed in the largest group (Group 1), while the isolates R1‐R6 (from Raphanus) were usually placed in a distinct but smaller group (Group 2). There were only approximately 90% identical nucleotides between the two groups. However, one isolate (R5) showed evidence of recombination in that the region between nucleotides 430 and 450, from the start of the coat protein gene and its 3′‐terminus, was a Group 1 type.  相似文献   

12.
Three rat BV13S1 alleles (T-cell receptor β-chain variable gene 13) were characterized by new BV13S1-allele specific monoclonal antibodies (18B1 and 17D5) and sequence analysis of expressed and genomic BV13S1. Two alleles were functional and designated BV13S1A1 present in strains LEW, BUF, PVG, and BV13S1A2 present in BN and WF. Their products differed by six amino acids, two of them in complementarity-determing region (CDR)1 and one in CDR2. A third nonfunctional allele, BV13S1A3P, was found in strains F344 and DA. Apart from a single nucleotide insertion, it was identical to BV13S1A2. All 12 rat strains tested showed association of TCRBC1 with BV8S2/4 alleles but not with the BV13S1 alleles, which may reflect a different gene order of the rat BV compared to mouse. BV13S1A1-encoded T-cell receptors (TCRs) which bind both monoclonal antibody (mAb) 18B1 and mAb 17D5 are over-represented in the CD4 lymphocyte subset. BV13S1A2-encoded TCRs which are stained by mAb 18B1 but not by mAb 17D5 show a slight CD8-biased expression. Preferential usage of BV13S1A1-positive TCRs by CD4 but not by CD8 cells in (LEW×WF)F1 hybrids and cosegregation of BV13SA1 and increased frequency of BV13S1 TCR-positive CD4 cells in a (LEW×BN)×BN backcross suggest structural differences of the two allelic products as the reason for their contrasting CD4/CD8 subset bias. Received: 6 October 1999 / Revised: 25 November 1999  相似文献   

13.
To evaluate the polymorphism and conservation of the major histocompatibility complex class Ib molecule Qa1 in wild mouse populations, we determined the nucleotide sequence of exons 1–3 of Qa1 of eight mouse haplotypes derived from wild mice, including Mus musculus domesticus, M. m. castaneus, M. m. bactrianus, and M. spretus, as well as two t haplotypes. Our data identify eight new alleles of Qa1. Taken together with previously published data on Qa1 among the common laboratory inbred strains, and in agreement with cytotoxic T-lymphocyte, serological, and biochemical data, these results further confirm the existence of two families of Qa1 molecules, Qa1a-like and Qa1b-like, and illuminate the extreme conservation of the peptide-binding region of these molecules, even across species.The wild mouse Qa1 nucleotide sequences are available from GenBank at accession numbers AF100695–703  相似文献   

14.
The por genes of the predominant serovars of Neisseria gonorrhoeae circulating in a high-frequency transmitter core group located in Nairobi, Kenya, were examined for nucleotide sequence polymorphism. The level of por gene diversity did not differ significantly between core group-derived gonococcal strains and gonococcal strains originating elsewhere. However, por mosaicism appeared to be more frequent among core group-derived strains, suggesting that recombination of different por sequences may be a important strategy by which N. gonorrhoeae generates por gene diversity within core group populations. Despite extensive sequence variability, por expressed by gonococcal isolates of different geographic origin exhibited conserved patterns of nucleotide change, suggesting that diversity among por alleles may also be finite.  相似文献   

15.
16.
Helicobacter pylori is a genetically diverse organism that is adapted for colonization of the human stomach. All strains contain a gene encoding a secreted, pore-forming toxin known as VacA. Genetic variation at this locus could be under strong selection as H. pylori adapts to the host immune response, colonizes new human hosts, or inhabits different host environments. Here, we analyze the molecular evolution of VacA. Phylogenetic reconstructions indicate the subdivision of VacA sequences into three main groups with distinct geographic distributions. Divergence of the three groups is principally due to positively selected sequence changes in the p55 domain, a central region required for binding of the toxin to host cells. Divergent amino acids map to surface-exposed sites in the p55 crystal structure. Comparative phylogenetic analyses of vacA sequences and housekeeping gene sequences indicate that vacA does not share the same evolutionary history as the core genome. Further, rooting the VacA tree with outgroup sequences from the close relative Helicobacter acinonychis reveals that the ancestry of VacA is different from the African origin that typifies the core genome. Finally, sequence analyses of the virulence determinant CagA reveal three main groups strikingly similar to the three groups of VacA sequences. Taken together, these results indicate that positive selection has shaped the phylogenetic structure of VacA and CagA, and each of these virulence determinants has evolved separately from the core genome.Helicobacter pylori is a Gram-negative bacterium that persistently colonizes the human stomach. H. pylori induces a gastric mucosal inflammatory response known as superficial gastritis and is a risk factor for the development of peptic ulcer disease, gastric adenocarcinoma, and gastric mucosa-associated lymphoid tissue (MALT) lymphoma (2, 43). H. pylori is present in about half of all humans throughout the world.H. pylori strains from unrelated humans exhibit a high level of genetic diversity (5, 44). The population structure of H. pylori is panmictic, and the rate of recombination in H. pylori is reported to be among the highest in the Eubacteria (17, 44). Multilocus sequence analysis of housekeeping genes has revealed the presence of at least nine different H. pylori populations or subpopulations that are localized to distinct geographic regions (12, 27, 31). Analysis of these sequences suggests that H. pylori has spread throughout the world concurrently with the major events of human dispersal, and thus H. pylori is potentially a useful marker for the geographic migrations of human populations (12).One of the important virulence determinants of H. pylori is a secreted toxin known as VacA. VacA is a pore-forming toxin that causes multiple alterations in human cells, including cell vacuolation, depolarization of membrane potential, alteration of mitochondrial membrane permeability, apoptosis, activation of mitogen-activated protein kinases, inhibition of antigen presentation, and inhibition of T-cell activation and proliferation (8, 10, 15). Secreted by an autotransporter (type Va) secretion mechanism, VacA is translated as a 140-kDa protoxin that undergoes N- and C-terminal cleavage during the secretion process to yield an N-terminal signal sequence, a mature 88-kDa secreted toxin known as p88, a small secreted peptide with no known function (termed secreted alpha peptide, or SAP) (7), and a C-terminal beta-barrel domain (41, 47) (Fig. (Fig.1A).1A). Two domains of p88 VacA, p33 and p55, have been identified based on partial proteolysis of p88 into fragments of 33 kDa and 55 kDa, respectively (47) (Fig. (Fig.1A).1A). The N-terminal p33 domain (residues 1 to 311) is involved in pore formation while the p55 domain (residues 312 to 821) contains one or more cell-binding domains (14, 48). The isolated p55 domain binds to host cells less avidly than does the full-length p88 protein, and in contrast to p88, the isolated p55 domain is not internalized by cells (18, 48). These observations suggest that sequences in both the p33 and p55 domains mediate VacA interactions with the surface of cells.Open in a separate windowFIG. 1.Analysis of VacA phylogeography. (A) The vacA gene encodes a 140-kDa protoxin, which undergoes cleavage to yield a signal sequence, a secreted 88-kDa toxin, a secreted alpha-peptide (SAP), and a C-terminal β-barrel domain. The mature 88-kDa VacA toxin contains two domains, designated p33 and p55. The midregion sequence that defines type m1 and m2 forms of VacA is located within p55. A 21-amino-acid insertion is present in m2 forms but not m1 forms of VacA. (B) Neighbor-joining phylogenetic tree of 100 amino acid sequences of VacA. Three major groups (designated groups 1 to 3) are evident. The chart shows the number of strains analyzed and characteristics of VacA protein sequences in each group of the tree. Group 1 comprises type m1 sequences mainly from non-Asian strains, group 2 comprises m1 sequences from Asian strains, and group 3 comprises m2 sequences from both Asian and non-Asian strains. See Fig. S1 in the supplemental material for a ladder-type version of this tree.All strains of H. pylori contain a chromosomal vacA gene, but individual strains differ considerably in levels of VacA activity (3, 8). Two studies analyzed vacA sequence encoding a fragment of the p33 domain and did not detect any recognizable phylogenetic structure (star or bush-type pattern), presumably due to the presence of extensive recombination (19, 44). Other studies analyzed different regions of VacA and detected polymorphisms that allow classification of vacA alleles into distinct families (designated s1/s2, i1/i2, and m1/m2) depending on the presence of signature sequences in different regions of VacA (3, 4, 39). Geographic differences have been detected within several of these vacA regions (22, 24, 29, 37, 51, 52, 55). In general, strains containing vacA alleles classified as s1, i1, or m1 have been associated with an increased risk of ulcer disease or gastric cancer compared to strains containing vacA alleles classified as s2, i2, or m2 (3, 13, 39).Another important H. pylori virulence factor is the secreted CagA effector protein. The cagA gene is localized within a 40-kb chromosomal region known as the cag pathogenicity island (PAI) (20). H. pylori strains expressing CagA are associated with a significantly increased risk for development of ulcer disease or gastric cancer compared to strains that lack the cagA gene (6). Upon entry into cells, CagA undergoes phosphorylation by host cell kinases and induces numerous alterations in cellular signaling, leading to the designation of CagA as a “bacterial oncoprotein” (20, 32).H. pylori strains that produce an active VacA protein (type s1 VacA) typically express CagA, and strains that produce inactive VacA proteins (type s2 VacA) typically lack the cagA gene (3). vacA and the cag PAI localize to distant sites on the H. pylori chromosome, and, therefore, the basis for this association has been unclear. Recently, several studies have reported that there are complex relationships between the cellular effects of VacA and CagA, whereby VacA can downregulate CagA''s effects on epithelial cells, or vice versa (1, 35, 46, 56). This functional interaction between VacA and CagA may represent a mechanism that allows H. pylori to minimize damage to gastric epithelial cells or minimize mucosal inflammation, thereby allowing it to persistently colonize the stomach.Although VacA is considered an important H. pylori virulence factor and hundreds of studies have classified H. pylori strains based on a vacA typing scheme, there has been very little effort to investigate the forces that drive vacA diversification, to analyze the evolutionary history of vacA, or to correlate vacA diversity with features of the VacA three-dimensional structure. Several important questions remain in studying the vacA gene: (i) Are the s1, i1, and m1 alleles (which are associated with an increased risk of gastroduodenal disease) more recently derived than the s2, i2, and m2 alleles? (ii) Are the geographic differences in vacA alleles driven by adaptive evolution or genetic drift? (iii) Does the evolutionary history of the vacA gene parallel the evolutionary history of the core genes used for MLST analysis, which are markers for ancient migrations of human populations?In the current study, we present a comprehensive analysis of the molecular evolution of vacA. Our analysis of VacA diversity indicates that VacA sequences are clustered into three main groups with distinct geographic distributions. By analyzing topological differences between vacA and housekeeping gene phylogenetic trees, we demonstrate that the vacA gene does not share the same evolutionary history as the core genome of H. pylori. We report that the evolution of VacA has been shaped by positive selection, and adaptive evolution is restricted to the p55 domain. Most of the sequence divergence corresponds to surface-exposed amino acids in the three-dimensional structure of the p55 domain. Finally, we note that there are similarities between the phylogenetic structure of the VacA and CagA trees, and we discuss the roles that positive selection pressures have played in the evolution of these two virulence determinants.  相似文献   

17.
The group A streptococcal emm-like genes, which encode the cell-surface M and M-like proteins, are divided into distinct mrp, emm and enn subdivisions and are clustered together in a region of the chromosome called the vir regulon. In order to understand the mechanisms involved in the evolution of emm-like genes, a 180bp fragment of the 5 variable region of the enn gene was characterized in 31 strains for which emm sequences and multilocus enzyme electrophoretic profiles have been previously determined. The results demonstrate that nucleotide polymorphisms at the enn locus are generated predominantly by point mutations and short deletions or insertions, and that variation among enn and emm genes has arisen by similar mechanisms. However, diversity at the enn locus is restricted in comparison to the emm locus. Moreover, there is strong evidence for intragenic recombination at the enn locus and the pattern of distribution of emm and enn alleles among strains suggests that these genes may be independently acquired by horizontal transfer and recombination from distinct donor strains, thereby generating a mosaic structure for the vir regulon. The results add to a growing body of evidence that horizontal gene transfer has played a major role in the evolution of Streptococcus pyogenes vir regulons.  相似文献   

18.
 At least 32 mostly single-member subfamilies of T-cell receptor alpha variable (TCRAV) genes have been described in humans. The AV1 subfamily is the largest, estimated by hybridization to contain as many as five members. However, a search of nucleotide sequence databases reveals a much greater number of unique sequences corresponding to this subfamily. In order to resolve this discrepancy between hybridization and nucleotide sequencing data, and to better understand the nature of variability among variable genes within a large subfamily, a genomic characterization of the AV1 subfamily in humans was carried out. Total genomic DNA, as well as isolated genomic clones spanning the TCRA region were screened for members of the AV1 subfamily by polymerase chain reaction (PCR) and nucleotide sequencing as well as by hybridization. A total of eight AV1 genes were identified and their nucleotide sequences were determined. Three of the sequences represent new genes. Based on structural features and the results of PCR screening of cDNA, none of these new genes appear to be functional. Several additional previously reported AV1 sequences were determined to represent alleles of AV1 genes, and simple PCR restriction digest assays were established for their detection. Use of each of the identified AV1 genes as hybridization probes failed to reveal any additional hybridizing bands. Thus the AV1genes represent the largest TCRAV subfamily with a maximum of eight members, several of which have common allelic forms. Received: 7 November 1996 / Revised: 5 December 1996  相似文献   

19.
Genetic recombination is considered to be a very frequent phenomenon among enteroviruses (Family Picornaviridae, Genus Enterovirus). However, the recombination patterns may differ between enterovirus species and between types within species. Enterovirus C (EV-C) species contains 21 types. In the capsid coding P1 region, the types of EV-C species cluster further into three sub-groups (designated here as A–C). In this study, the recombination pattern of EV-C species sub-group B that contains types CVA-21, CVA-24, EV-C95, EV-C96 and EV-C99 was determined using partial 5′UTR and VP1 sequences of enterovirus strains isolated during poliovirus surveillance and previously published complete genome sequences. Several inter-typic recombination events were detected. Furthermore, the analyses suggested that inter-typic recombination events have occurred mainly within the distinct sub-groups of EV-C species. Only sporadic recombination events between EV-C species sub-group B and other EV-C sub-groups were detected. In addition, strict recombination barriers were inferred for CVA-21 genotype C and CVA-24 variant strains. These results suggest that the frequency of inter-typic recombinations, even within species, may depend on the phylogenetic position of the given viruses.  相似文献   

20.
Hypervariability of intronic simple (gt)n(ga)m repeats in HLA-DRB genes   总被引:2,自引:2,他引:0  
We have investigated the extent of DNA variability in intronic simple (gt)n(ga)m repeat sequences and correlated this to sequence polymorphisms in the flanking exon 2 of HLA-DRB genes. The polymerase chain reaction (PCR) was used to amplify a DNA fragment containing exon 2 and the repeat region of intron 2. The PCR products were separated on sequencing gels in order to demonstrate length hypervariability of the (gt)n(ga)m repeats. In a parallel experiment, the PCR products were cloned and sequenced (each exon 2 plus adjacent simple repeats) to characterize the simple repeats in relation to the HLA-DRB sequences. In a panel of 25 DRB1, DRB4, and DRB5 alleles new sequences were not detected. Restriction fragment length polymorphism (RFLP) subtyping of serologically defined haplotypes corresponds to translated DNA sequences in 85% of the cases, the exceptions involving unusual DR/DQ combinations. Many identical DRB1 alleles can be distinguished on the basis of their adjacent simple repeats. We found group-specific organization of the repeats: the DRw52 supergroup repeats differ from those of DRB1*0101, DRB4*0101, and DRB5*0101 alleles and from those of pseudogenes. Finally, we amplified baboon DNA and found a DRB allele with extensive similarity to DRB1 sequences of the DRw52 supergroup. The simple repeat of the baboon gene, however, resembles that of human pseudogenes. In addition to further subtyping, the parallel study of polymorphic protein and hypervariable DNA alleles may allow conclusions to be drawn on the relationships between the DRB genes and perhaps also on the theory of trans-species evolution.The nucleotide sequence data reported in this paper have been submitted to the GenBank nucleotide sequence database and have been assigned the accession number M 34258.  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