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.
We have previously used NotI as the marker enzyme (recognizing GCGGCCGC) in a genome scanning approach for detection of mutations induced in mouse spermatogonia and estimated the mutation induction rate as about 0.7 x 10(-5) per locus per Gy. To see whether different parts of the genome have different sensitivities for mutation induction, we used AflII (recognizing CTTAAG) as the marker enzyme in the present study. After the screening of 1,120 spots in each mouse offspring, we found five mutations among 92,655 spots from the unirradiated paternal genome, five mutations among 218,411 spots from the unirradiated maternal genome, and 13 mutations among 92,789 spots from 5 Gy-exposed paternal genome. Among the 23 mutations, 11 involved mouse satellite DNA sequences (AT-rich), and the remaining 12 mutations also involved AT-rich but non-satellite sequences. Both types of sequences were found as multiple, similar-sequence blocks in the genome. Counting each member of cluster mutations separately and excluding results on one hypermutable spot, the spontaneous mutation rates were estimated as 3.2 (+/- 1.9) x 10(-5) and 2.3 (+/- 1.0) x 10(-5) per locus per generation in the male and female genomes, respectively, and the mutation induction rate as 1.1 (+/- 1.2) x 10(-5) per locus per Gy. The induction rate would be reduced to 0.9 x 10(-5) per locus per Gy if satellite sequence mutations were excluded from this analysis. The results indicate that mutation induction rates do not largely differ between GC-rich and AT-rich regions: 1 x 10(-5) per locus per Gy or less, which is close to 1.08 x 10(-5) per locus per Gy, the current estimate for the mean mutation induction rate in mice.  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.
A per-generation somatic mutation rate for microsatellites was estimated in western redcedar (Thuja plicata, Donn ex D. Don.: Cupressaceae). A total of 80 trees representative of the average size and age of reproductive trees were sampled in four natural populations in southwestern British Columbia. Samples of bulked haploid megagametophytes were collected from two or three positions on each tree, assuming that the collections were far enough apart that the same mutant sector was not sampled twice. All samples were genotyped at eight microsatellite loci. A single mutation corresponding to a stepwise increase in one dinucleotide repeat was detected. The estimated mutation rate for microsatellites was 6.3 x 10(-4) mutations per locus per generation (or 3.1 x 10(-4) per allele per generation), with a 95% confidence interval of 3.0 x 10(-5) to 4.0 x 10(-3) mutations per locus. Somatic mutations can contribute to a greater mutational load in trees, as compared to shorter lived plants, and genotypic mosaics within an individual have important implications for plant defense strategies and plant evolution.  相似文献   

4.
Tanaka MM  Bergstrom CT  Levin BR 《Genetics》2003,164(3):843-854
Recent studies have found high frequencies of bacteria with increased genomic rates of mutation in both clinical and laboratory populations. These observations may seem surprising in light of earlier experimental and theoretical studies. Mutator genes (genes that elevate the genomic mutation rate) are likely to induce deleterious mutations and thus suffer an indirect selective disadvantage; at the same time, bacteria carrying them can increase in frequency only by generating beneficial mutations at other loci. When clones carrying mutator genes are rare, however, these beneficial mutations are far more likely to arise in members of the much larger nonmutator population. How then can mutators become prevalent? To address this question, we develop a model of the population dynamics of bacteria confronted with ever-changing environments. Using analytical and simulation procedures, we explore the process by which initially rare mutator alleles can rise in frequency. We demonstrate that subsequent to a shift in environmental conditions, there will be relatively long periods of time during which the mutator subpopulation can produce a beneficial mutation before the ancestral subpopulations are eliminated. If the beneficial mutation arises early enough, the overall frequency of mutators will climb to a point higher than when the process began. The probability of producing a subsequent beneficial mutation will then also increase. In this manner, mutators can increase in frequency over successive selective sweeps. We discuss the implications and predictions of these theoretical results in relation to antibiotic resistance and the evolution of mutation rates.  相似文献   

5.
In a recent study, we reported that the combined average mutation rate of 10 di-, 6 tri-, and 8 tetranucleotide repeats in Drosophila melanogaster was 6.3 x 10(-6) mutations per locus per generation, a rate substantially below that of microsatellite repeat units in mammals studied to date (range = 10(-2)-10(-5) per locus per generation). To obtain a more precise estimate of mutation rate for dinucleotide repeat motifs alone, we assayed 39 new dinucleotide repeat microsatellite loci in the mutation accumulation lines from our earlier study. Our estimate of mutation rate for a total of 49 dinucleotide repeats is 9.3 x 10(-6) per locus per generation, only slightly higher than the estimate from our earlier study. We also estimated the relative difference in microsatellite mutation rate among di-, tri-, and tetranucleotide repeats in the genome of D. melanogaster using a method based on population variation, and we found that tri- and tetranucleotide repeats mutate at rates 6.4 and 8.4 times slower than that of dinucleotide repeats, respectively. The slower mutation rates of tri- and tetranucleotide repeats appear to be associated with a relatively short repeat unit length of these repeat motifs in the genome of D. melanogaster. A positive correlation between repeat unit length and allelic variation suggests that mutation rate increases as the repeat unit lengths of microsatellites increase.   相似文献   

6.
Estimate of the mutation rate per nucleotide in humans   总被引:41,自引:0,他引:41  
Nachman MW  Crowell SL 《Genetics》2000,156(1):297-304
Many previous estimates of the mutation rate in humans have relied on screens of visible mutants. We investigated the rate and pattern of mutations at the nucleotide level by comparing pseudogenes in humans and chimpanzees to (i) provide an estimate of the average mutation rate per nucleotide, (ii) assess heterogeneity of mutation rate at different sites and for different types of mutations, (iii) test the hypothesis that the X chromosome has a lower mutation rate than autosomes, and (iv) estimate the deleterious mutation rate. Eighteen processed pseudogenes were sequenced, including 12 on autosomes and 6 on the X chromosome. The average mutation rate was estimated to be approximately 2.5 x 10(-8) mutations per nucleotide site or 175 mutations per diploid genome per generation. Rates of mutation for both transitions and transversions at CpG dinucleotides are one order of magnitude higher than mutation rates at other sites. Single nucleotide substitutions are 10 times more frequent than length mutations. Comparison of rates of evolution for X-linked and autosomal pseudogenes suggests that the male mutation rate is 4 times the female mutation rate, but provides no evidence for a reduction in mutation rate that is specific to the X chromosome. Using conservative calculations of the proportion of the genome subject to purifying selection, we estimate that the genomic deleterious mutation rate (U) is at least 3. This high rate is difficult to reconcile with multiplicative fitness effects of individual mutations and suggests that synergistic epistasis among harmful mutations may be common.  相似文献   

7.
Transabdominal X-rays are a risk factor for childhood leukemia, and X-ray exposure of mouse fetuses has led to increases in both mutations and initiated tumors in offspring. However, fetal sensitivity and dose-response characteristics with regard to transplacental mutagenesis by X-rays have never been quantified. In the current experiment, pregnant Syrian hamsters at day 12 of gestation were irradiated with 300-kV X-rays. Twenty-four hours later, the fetuses were removed and their cells were allowed a 5 day expression time in culture. They were then seeded for colony formation and also for mutation selection by 6-thioguanine (6-TG). Mutation frequency was linear over the entire dose range, 10-600 R. The average induced 6-TG mutant frequency was 4.7 x 10(-7) per R. These results suggest that fetal cells are highly sensitive to induction of mutations by X-rays, and that a no-effect threshold is not likely. The 10 R dose caused a 25-fold increase in mutation frequency over the historical control, 45 x 10(-7) versus 1.8 x 10(-7), an increase per R of 2.5-fold. Increased risk of childhood cancer related to obstetrical transabdominal X-ray has also been estimated at 2.5-fold per R. Thus, our results are consistent with mutation contributing to this effect.  相似文献   

8.
A subclone of a human diploid lymphoblastoid cell line, TK-6, with consistently high cloning efficiency has been used to estimate the rates of somatic mutations on the basis of protein variation detected by two-dimensional polyacrylamide gel electrophoresis. A panel of 267 polypeptide spots per gel was screened, representing the products of approximately 263 unselected loci. The rate of human somatic mutation in vitro was estimated by measuring the proportion of protein variants among cell clones isolated at various times during continuous exponential growth of a TK-6 cell population. Three mutants of spontaneous origin were observed, giving an estimated spontaneous rate of 6 x 10(-8) electrophoretic mutations per allele per cell generation (i.e., 1.2 x 10(-7) per locus per cell generation). Following treatment of cells with N-ethyl-N-nitrosourea, a total of 74 confirmed variants at 54 loci were identified among 1143 clones analyzed (approximately 601,000 allele tests). The induced variants include 65 electromorphs which exhibit altered isoelectric charge and/or apparent molecular weight and nine nullimorphs for each of which a gene product was not detected at its usual location on the gel. The induced frequency for these 65 structural gene mutants is 1.1 x 10(-4) per allele. An excess of structural gene mutations at ten known polymorphic loci and repeat mutations at these and other loci suggest nonrandomness of mutation in human somatic cells. Nullimorphs occurring at three heterozygous loci in TK-6 cells may be caused by genetic processes other than structural gene mutation.  相似文献   

9.
Within recent years, microsatellite have become one of the most powerful genetic markers in biology. For several mammalian species, microsatellite mutation rates have been estimated on the order of 10(- 3)-10(-5). A recent study, however, demonstrated mutation rates in Drosophila melanogaster of at least one order of magnitude lower than those in mammals. To further test this result, we examined mutation rates of different microsatellite loci using a larger sample size. We screened 24 microsatellite loci in 119 D. melanogaster lines maintained for approximately 250 generations and detected 9 microsatellite mutations. The average mutation rate of 6.3 x 10(-6) is identical to the mutation rate from a previous study. Most interestingly, all nine mutations occurred at the same allele of one locus (DROYANETSB). This hypermutable allele has 28 dinucleotide repeats and is among the longest microsatellite reported in D. melanogaster. The allele-specific mutation rate of 3.0 x 10(-4) per generation is within the range of mammalian mutation rates. Future microsatellite analyses will have to account for the dramatic differences in allele-specific mutation rates.   相似文献   

10.
The compost worm Eisenia fetida is routinely used in ecotoxicological studies. A standard assay to assess genetic damage in this species would be extremely valuable. Since mitochondrial DNA (mtDNA) is known to exhibit an increased mutation rate following exposure to ionising radiation we assessed the validity of a mtDNA-based assay for measuring increases in mutation rate in laboratory-irradiated compost worms. To this end the mutation frequency in the mtDNA of the compost worm E. fetida was quantified following in vivo gamma-irradiation of adult worms in three dose groups. Five adult worms exposed to 1.4 mGy/h for 55 days (total dose 1.85 Gy), five adult worms exposed to 8.5 mGy/h for 55 days (total dose 11.22 Gy) and five adult control worms were used to assess the effect of irradiation on mtDNA mutation induction. DNA samples extracted from irradiated adult worms were used in high-fidelity PCR of a 486 bp region of mtDNA spanning the ATPase 8 gene, chosen for its high spontaneous mutation rate. PCR products were cloned and sequenced to identify mutations, with 89-102 clones successfully sequenced per individual. A significant elevation in mtDNA mutation frequency (p=0.032) was seen in worms exposed at the higher dose rate (8.5 mGy/h, total dose 11.22 Gy; mutation frequency 27.98+/-4.85 x 10(-5)mutations/bp) in comparison to controls (mutation frequency 12.68+/-3.06 x 10(-5)mutations/bp), but no elevation in mutation frequency (p=0.764) was seen for the lower dose rate (1.4 mGy/h, total dose 1.85 Gy; mutation frequency 13.74+/-1.29 x 10(-5)mutations/bp) compared with controls. This indicates that although the technique has the potential to detect an elevation in mutation frequency, it does not have sufficient sensitivity at the doses likely to be encountered in environmental monitoring scenarios.  相似文献   

11.
Lang GI  Murray AW 《Genetics》2008,178(1):67-82
Although mutation rates are a key determinant of the rate of evolution they are difficult to measure precisely and global mutations rates (mutations per genome per generation) are often extrapolated from the per-base-pair mutation rate assuming that mutation rate is uniform across the genome. Using budding yeast, we describe an improved method for the accurate calculation of mutation rates based on the fluctuation assay. Our analysis suggests that the per-base-pair mutation rates at two genes differ significantly (3.80x10(-10) at URA3 and 6.44x10(-10) at CAN1) and we propose a definition for the effective target size of genes (the probability that a mutation inactivates the gene) that acknowledges that the mutation rate is nonuniform across the genome.  相似文献   

12.
Notley-McRobb L  Seeto S  Ferenci T 《Genetics》2002,162(3):1055-1062
The kinetics of mutator sweeps was followed in two independent populations of Escherichia coli grown for up to 350 generations in glucose-limited continuous culture. A rapid elevation of mutation rates was observed in both populations within 120-150 generations, as was apparent from major increases in the proportion of the populations with unselected mutations in fhuA. The increase in mutation rates was due to sweeps by mutY mutators. In both cultures, the enrichment of mutators resulted from hitchhiking with identified beneficial mutations increasing fitness under glucose limitation; mutY hitchhiked with mgl mutations in one culture and ptsG in the other. In both cases, mutators were enriched to constitute close to 100% of the population before a periodic selection event reduced the frequency of unselected mutations and mutators in the cultures. The high proportion of mutators persisted for 150 generations in one population but began to be eliminated within 50 generations in the other. The persistence of mutator, as well as experimental data showing that mutY bacteria were as fit as near-isogenic mutY(+) bacteria in competition experiments, suggest that mutator load by deleterious mutations did not explain the rapidly diminishing proportion of mutators in the populations. The nonmutators sweeping out mutators were also unlikely to have arisen by reversion or antimutator mutations; the mutY mutations were major deletions in each case and the bacteria sweeping out mutators contained intact mutY. By following mgl allele frequencies in one population, we discovered that mutators were outcompeted by bacteria that had rare mgl mutations previously as well as additional beneficial mutation(s). The pattern of appearance of mutY, but not its elimination, conforms to current models of mutator sweeps in bacterial populations. A mutator with a narrow mutational spectrum like mutY may be lost if the requirement for beneficial mutations is for changes other than GC --> TA transversions. Alternatively, epistatic interactions between mutator mutation and beneficial mutations need to be postulated to explain mutator elimination.  相似文献   

13.
Joseph SB  Hall DW 《Genetics》2004,168(4):1817-1825
We performed a 1012-generation mutation-accumulation (MA) experiment in the yeast, Saccharomyces cerevisiae. The MA lines exhibited a significant reduction in mean fitness and a significant increase in variance in fitness. We found that 5.75% of the fitness-altering mutations accumulated were beneficial. This finding contradicts the widely held belief that nearly all fitness-altering mutations are deleterious. The mutation rate was estimated as 6.3 x 10(-5) mutations per haploid genome per generation and the average heterozygous fitness effect of a mutation as 0.061. These estimates are compatible with previous estimates in yeast.  相似文献   

14.
Nilsson AI  Kugelberg E  Berg OG  Andersson DI 《Genetics》2004,168(3):1119-1130
Experimental evolution is a powerful approach to study the dynamics and mechanisms of bacterial niche specialization. By serial passage in mice, we evolved 18 independent lineages of Salmonella typhimurium LT2 and examined the rate and extent of adaptation to a mainly reticuloendothelial host environment. Bacterial mutation rates and population sizes were varied by using wild-type and DNA repair-defective mutator (mutS) strains with normal and high mutation rates, respectively, and by varying the number of bacteria intraperitoneally injected into mice. After <200 generations of adaptation all lineages showed an increased fitness as measured by a faster growth rate in mice (selection coefficients 0.11-0.58). Using a generally applicable mathematical model we calculated the adaptive mutation rate for the wild-type bacterium to be >10(-6)/cell/generation, suggesting that the majority of adaptive mutations are not simple point mutations. For the mutator lineages, adaptation to mice was associated with a loss of fitness in secondary environments as seen by a reduced metabolic capability. During adaptation there was no indication that a high mutation rate was counterselected. These data show that S. typhimurium can rapidly and extensively increase its fitness in mice but this niche specialization is, at least in mutators, associated with a cost.  相似文献   

15.
16.
We estimated the rates per base per generation of specific types of mutations, using our direct estimate of the overall mutation rate for hemophilia B and information on the mutations present in the United Kingdom's population as well as those reported year by year in the hemophilia B world database. These rates are as follows: transitions at CpG sites 9.7x10-8, other transitions 7.3x10-9, transversions at CpG sites 5.4x10-9, other transversions 6.9x10-9, and small deletions/insertions causing frameshifts 3.2x10-10. By taking into account the ratio of male to female mutation rates, the above figures were converted into rates appropriate for autosomal DNA-namely, 1.3x10-7, 9.9x10-9, 7.3x10-9, 9.4x10-9, 6.5x10-10, where the latter is the rate for all small deletion/insertion events. Mutation rates were also independently estimated from the sequence divergence observed in randomly chosen sequences from the human and chimpanzee X and Y chromosomes. These estimates were highly compatible with those obtained from hemophilia B and showed higher mutation rates in the male, but they showed no evidence for a significant excess of transitions at CpG sites in the spectrum of Y-sequence divergence relative to that of X-chromosome divergence. Our data suggest an overall mutation rate of 2.14x10-8 per base per generation, or 128 mutations per human zygote. Since the effective target for hemophilia B mutations is only 1.05% of the factor IX gene, the rate of detrimental mutations, per human zygote, suggested by the hemophilia data is approximately 1.3.  相似文献   

17.
Error-prone polymerase chain reaction (PCR) is widely used to introduce point mutations during in vitro evolution experiments. Accurate estimation of the mutation rate during error-prone PCR is important in studying the diversity of error-prone PCR product. Although many methods for estimating the mutation rate during PCR are available, all the existing methods depend on the assumption that the mutation rate is low and mutations occur at different places whenever they occur. The available methods may not be applicable to estimate the mutation rate during error-prone PCR. We develop a mathematical model for error-prone PCR and present methods to estimate the mutation rate during error-prone PCR without assuming low mutation rate. We also develop a computer program to simulate error-prone PCR. Using the program, we compare the newly developed methods with two other methods. We show that when the mutation rate is relatively low (< 10(-3) per base per PCR cycle), the newly developed methods give roughly the same results as previous methods. When the mutation rate is relatively high (> 5 x 10(-3) per base per PCR cycle, the mutation rate for most error-prone PCR experiments), the previous methods underestimate the mutation rate and the newly developed methods approximate the true mutation rate.  相似文献   

18.
19.
In this study, we measured the frequency of revertants of a cytopathic strain of the duck hepatitis B virus that bears a single nucleotide substitution in the pre-S envelope protein open reading frame, resulting in the amino acid substitution G133E. Cytopathic virus mixed with known amounts of a genetically marked wild-type virus was injected into ducklings. Virus outgrowth was accompanied by a coselection of wild-type and spontaneous revertants during recovery of the ducklings from the acute liver injury caused by death of the G133E-infected cells. The frequency of individual revertants in the selected noncytopathic virus population was estimated by determining the ratio of each revertant to the wild-type virus. Spontaneous revertants were found to be present at frequencies of 1 x 10(-5) to 6 x 10(-5) per G133E genome inoculated. A mathematical model was used to estimate that the mutation rate was 0.8 x 10(-5) to 4.5 x 10(-5) per nucleotide per generation.  相似文献   

20.
Seo TK  Thorne JL  Hasegawa M  Kishino H 《Genetics》2002,160(4):1283-1293
Using pseudomaximum-likelihood approaches to phylogenetic inference and coalescent theory, we develop a computationally tractable method of estimating effective population size from serially sampled viral data. We show that the variance of the maximum-likelihood estimator of effective population size depends on the serial sampling design only because internal node times on a coalescent genealogy can be better estimated with some designs than with others. Given the internal node times and the number of sequences sampled, the variance of the maximum-likelihood estimator is independent of the serial sampling design. We then estimate the effective size of the HIV-1 population within nine hosts. If we assume that the mutation rate is 2.5 x 10(-5) substitutions/generation and is the same in all patients, estimated generation lengths vary from 0.73 to 2.43 days/generation and the mean (1.47) is similar to the generation lengths estimated by other researchers. If we assume that generation length is 1.47 days and is the same in all patients, mutation rate estimates vary from 1.52 x 10(-5) to 5.02 x 10(-5). Our results indicate that effective viral population size and evolutionary rate per year are negatively correlated among HIV-1 patients.  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