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.
2.
The principle of the treatment condition routinely used in Salmonella typhimurium is to allow the cells to divide in the presence of the chemical being tested; only the revertants will be able to form visible colonies (softagar procedure). In Saccharomyces cerevisiae, the routinely used procedure is to treat the cells in liquid non-nutrient medium under non-growing conditions (non-nutrient procedure). We compared mutation induction under both experimental conditions using S. cerevisiae; we also compared the mutagenic response of the two microorganisms to six compounds; two nitrofuran derivatives, AF-2 and SQ18,506, three hair dye components, 1,2-diamino-4-nitrobenzene, 1,4-diaminoanthraquinone, and methyl violet, as well as ethyl methanesulfonate. Of the six compounds tested in S. cerevisiae strain XV185-14C, only ethyl methanesulfonate was mutagenic under both experimental conditions. The two nitrofuran derivatives, AF-2 and SQ18,506, induced mutations in S. cerevisiae when the non-nutrient procedure was employed. None of the three hair dyes tested was mutagenic in S. cerevisiae. However, the results obtained with Salmonella typhimurium indicate that the hair dye 1,2-diamino-4-nitrobenzene is a mutagen, confirming the earlier study by Ames et al. [2]. Among the other five compounds tested in Salmonella typhimurium, the base-substitution-detecting strain TA100 responded to one concentration of AF-2, and EMS was mutagenic in strains TA1535, TA100 and TA1537.  相似文献   

3.
The spectrum of mutations induced by the carcinogen N-2-acetylaminofluorene (AAF) was analysed in Saccharomyces cerevisiae using a forward mutation assay, namely the inactivation of the URA3 gene. The URA3 gene, carried on a yeast/bacterial shuttle vector, was randomly modified in vitro using N-acetoxy-N-2-acetylaminofluorene (N-AcO-AAF) as a model reactive metabolite of the carcinogen AAF. The binding spectrum of AAF to the URA3 gene was determined and found to be essentially random, as all guanine residues reacted about equally well with N-AcO-AAF. Independent Ura? mutants were selected in vivo after transformation of the modified plasmid into a ura3Δ yeast strain. Plasmid survival decreased as a function of AAF modification, leading to one lethal hit (37% relative survival) for an average of ≈ 50 AAF adducts per plasmid molecule. At this level of modification the mutation frequency was equal to ≈ 70 × 10?4, i.e. ≈ 50-fold above the background mutation frequency. UV irradiation of the yeast cells did not further stimulate the mutagenic response, indicating the lack of an SOS-like mutagenic response in yeast. Sequence analysis of the URA3 mutants revealed ≈ 48% frameshifts, 44% base substitutions and ≈ 8 % complex events. While most base substitutions (74%) were found to be targeted at G residues where AAF is known to form covalent C8 adducts, frameshift mutations were observed at GC base pairs in only≈ 24% of cases. Indeed, more than 60% of frameshift events occurred at sequences such as 5′-(A/T)nG-3′ where a short (n = 2 or 3) monotonous run of As or Ts is located on the 5' side of a guanine residue. We refer to these mutations as semi-targeted events and present a potential mechanism that explains their occurrence.  相似文献   

4.
Summary The aroC321 allele permits positive selection for the detection of a large genetic duplication that arises in the Salmonella typhimurium chromosome by homologous recombination. Strains that contain both aroC321 and the hisC3076 allele were constructed so that the induction of genetic duplications and frameshift mutations in a run of GC base pairs could be studied simultaneously by selecting for tryptophan and histidine prototrophy, respectively. Using these strains, we examined the ability of 9-aminoacridine, quinacrine, four acridine mustards (ICR-170, ICR-191, ICR-372, and quinacrine mustard) and the nitroacridine Entozon to induce genetic duplications and frameshift mutations. Although all these compounds induce reversion of hisC3076, only the four mustards and Entozon are effective as inducers of genetic duplications under identical treatment conditions. The induction of genetic duplications by acridine mustards, like the toxic and mutagenic effects of these compounds, is enhanced by a deficiency for excision repair caused by a deletion through the uvrB gene. The ineffectiveness of 9-aminoacridine and quinacrine in the test for genetic duplications indicates that simple intercalation is sufficient for the mutagenic effect measured with the hisC3076 allele but that the induction of duplications by the acridine mustards and Entozon requires covalent binding of the chemical to DNA.  相似文献   

5.
We have used the lacZ reversion assay to study the mutation spectra induced by the Escherichia coli chromosomal umuDC operon and of its two plasmid-borne analogues impCAB and mucAB following exposure of cells to UV light and methyl methane-sulfonate (MMS). We have shown that the impCAB, mucAB and umuDC operons all produce a similar response to UV light which results almost exclusively in AT → GC transitions. However, we found that the three operons produced different responses to alkylating agents. We found that with MMS the chromosomal umuDC operon produced almost exclusively AT → GC transitions, whilst both mucAB and impCAB produced predominantly transversions. In the case of the impCAB operon the mutation spectrum contained more AT → TA than GC → TA transversions; this balance was reversed with mucAB. The effect of the copy number of the error-prone DNA repair operons upon the mutagenic spectra was also studied. The results obtained suggest that the copy number of the imp operon does not greatly affect the specificity of base substitutions observed. However, an increase in the copy number of the umuDC operon greatly affected the specificity of base substitution, such that virtually no transitions were produced and the spectrum was dominated by GC/AT → TA transversions. It appears that the three error-prone DNA repair operons impCAB, mucAB and umuDC, despite showing strong structural and functional homologies, can display major differences in the spectrum of base changes induced during mutagenesis. We propose that the type of misincorporation/chain extension which DNA polymerase III is allowed to synthesize on a damaged DNA template is extremely sensitive to both the amount and type of error-prone repair proteins present. The modulation of these events by the different proteins can result in widely different mutagenic changes in the repaired DNA.  相似文献   

6.
Translesion DNA synthesis (TLS) is a fundamental damage bypass pathway that utilises specialised polymerases with relaxed template specificity to achieve replication through damaged DNA. Misinsertions by low fidelity TLS polymerases may introduce additional mutations on undamaged DNA near the original lesion site, which we termed collateral mutations. In this study, we used whole genome sequencing datasets of chicken DT40 and several human cell lines to obtain evidence for collateral mutagenesis in higher eukaryotes. We found that cisplatin and UVC radiation frequently induce close mutation pairs within 25 base pairs that consist of an adduct-associated primary and a downstream collateral mutation, and genetically linked their formation to TLS activity involving PCNA ubiquitylation and polymerase κ. PCNA ubiquitylation was also indispensable for close mutation pairs observed amongst spontaneously arising base substitutions in cell lines with disrupted homologous recombination. Collateral mutation pairs were also found in melanoma genomes with evidence of UV exposure. We showed that collateral mutations frequently copy the upstream base, and extracted a base substitution signature that describes collateral mutagenesis in the presented dataset regardless of the primary mutagenic process. Using this mutation signature, we showed that collateral mutagenesis creates approximately 10–20% of non-paired substitutions as well, underscoring the importance of the process.  相似文献   

7.
The fidelity of initiator tRNA (i-tRNA) selection in the ribosomal P-site is a key step in translation initiation. The highly conserved three consecutive G:C base pairs (3GC pairs) in the i-tRNA anticodon stem play a crucial role in its selective binding in the P-site. Mutations in the 3GC pairs (3GC mutant) render the i-tRNA inactive in initiation. Here, we show that a mutation (E265K) in the unique C-terminal tail domain of RluD, a large ribosomal subunit pseudouridine synthase, results in compromised fidelity of initiation and allows initiation with the 3GC mutant i-tRNA. RluD modifies the uridine residues in H69 to pseudouridines. However, the role of its C-terminal tail domain remained unknown. The E265K mutation does not diminish the pseudouridine synthase activity of RluD, or the growth phenotype of Escherichia coli, or cause any detectable defects in the ribosomal assembly in our assays. However, in our in vivo analyses, we observed that the E265K mutation resulted in increased retention of the ribosome binding factor A (RbfA) on 30S suggesting a new role of RluD in contributing to RbfA release, a function which may be attributed to its (RluD) C-terminal tail domain. The studies also reveal that deficiency of RbfA release from 30S compromises the fidelity of i-tRNA selection in the ribosomal P-site.  相似文献   

8.
Mutagens were detected in the urine of rats following topical application of two commercial oxidative-type hair dye preparations. The test system used was induction of back mutation with the bacterial tester strain TA1538, a histidine-dependent mutant of Salmonella typhimurium. Various quantities of dye were applied to the shortened hair on the backs of the test animals. The dye was allowed to remain on the hair for 20 min after application and was then removed by shampooing and thorough rinsing. Maximal levels of mutagenic activity occurred with urine collected during the first 24 h following dye application, and a dose—response was observed when increasing volumes of mutagenic urine were tested.Mutagens were detected in rat urine after intraperitoneal injection, and also after topical application of 4-nitro-o-phenylenediamine, one of the constituents of the hair-dye preparations.  相似文献   

9.
140 Organophosphorus compounds (OP's) have been tested for mutagenic activity in bacteria, principally by using two specially constructed sets of tester strains of the bacteria Salmonella typhimurium and Escherichia coli. It was found that 20% gave positive mutagenic responses and that this group of chemicals produce base substitutions rather than frame-shift mutations. In most cases the DNA repair genes exrA+ and recA+ were for mutagenic activity.Seven compounds were further tested in Drosophila melanogaster for the ability to induce recessive lethal mutations. In some of these cases the doses administered to the flies had to be very low due to the highly toxic nature of the compounds. To overcome this problem, the accumulation of recessive lethal mutations was measured in populations which were continually exposed to the compounds over a period of some 18 months. During this time the populations developed increased resistance to the compound and so the dose administered could gradually be increased. Six of the compounds were mutagenic.Of the compounds tested in both systems, those showing mutagenic activity in bacteria were also mutaganic in Drosophila, those mutagenic in bacteria were not mutagenic in Drosophila.  相似文献   

10.
Spontaneous, 2-aminopurine- and 5-bromouracil-induced mutations at six rII nonsense codons were studied in phage T4 strains possessing wild-type and mutant gene 43 alleles. The mutation pathways studied included interconversions and reversions of nonsense codons. The tsCB87 allele, which specifies an antimutator DNA polymerase, reduced base-analogue-induced mutation frequencies along all pathways. However, GC base pairs were less affected than AT base pairs. The frequency of spontaneous UAA→UAG conversions was also reduced by tsCB87, but that of spontaneous UAA→UGA conversions was often increased. Mutation in the presence of the mutator allele tsL56 was increased along all pathways, with no preference for either AT or GC base pairs. Mutation frequencies in the presence of the two mutant DNA polymerases were highly variable. A strong correlation was found between 2-aminopurine-induced mutation frequencies in ts+ and tsCB87 phage along the reversion and UAA→UAG (but not UAA→UGA) pathways.  相似文献   

11.
Mutational fingerprints of aging   总被引:8,自引:0,他引:8       下载免费PDF全文
Using a lacZ plasmid transgenic mouse model, spectra of spontaneous point mutations were determined in brain, heart, liver, spleen and small intestine in young and old mice. While similar at a young age, the mutation spectra among these organs were significantly different in old age. In brain and heart G:C→A:T transitions at CpG sites were the predominant mutation, suggesting that oxidative damage is not a major mutagenic event in these tissues. Other base changes, especially those affecting A:T base pairs, positively correlated with increasing proliferative activity of the different tissues. A relatively high percentage of base changes at A:T base pairs and compound mutants were found in both spleen and spontaneous lymphoma, suggesting a possible role of the hypermutation process in splenocytes in carcinogenesis. The similar mutant spectra observed at a young age may reflect a common mutation mechanism for all tissues that could be driven by the rapid cell division that takes place during development. However, the spectra of the young tissues did not resemble that of the most proliferative aged tissue, implying that replicative history per se is not the underlying causal factor of age-related organ-specific differences in mutation spectra. Rather, differences in organ function, possibly in association with replicative history, may explain the divergence in mutation spectra during aging.  相似文献   

12.
The interaction of several 3,6-diaminoacridines with DNAs of various base composition has been studied by steady-state and transient fluorescence measurements. The acridine dyes employed are of the following two classes: class I - proflavine, acriflavine and 10-benzyl proflavine; class II - acridine yellow, 10-methyl acridine yellow and benzoflavine. It is found that the fluorescence decay kinetics follows a single-exponential decay law for free dye and the poly[d(A-T)]-dye complex, while that of the dye bound to DNA obeys a two-exponential decay law. The long lifetime (tau 1) for each complex is almost the same as the lifetime for the poly[d(A-T)]-dye complex, and the amplitude alpha 1 decreases with increasing GC content of DNA. The fluorescence quantum yields (phi F) of dye upon binding to DNA decrease with increasing GC content; the phi F values for class I are nearly zero when bound to poly(dG) X poly(dC), but those for class II are not zero. This is in harmony with the finding that GMP almost completely quenches the fluorescence for class I, whereas a weak fluorescence arises from the GMP-dye complex for class II. The fluorescence spectra of the DNA-dye complexes gradually shift toward longer wavelengths with increasing GC content. In this connection, the fluorescence decay parameters show a dependence on the emission wavelength; alpha 1 decreases with an increase in the emission wavelength. In view of these results, it is proposed that the decay behavior of the DNA-dye complexes has its origin in the heterogeneity of the emitting sites; the long lifetime tau 1 results from the dye bound to AT-AT sites, while the short lifetime tau 2 is attributable to the dye bound in the vicinity of GC pairs. Since GC pairs almost completely quench the fluorescence for class I, partly intercalated or externally bound dye molecules may play an important role in the component tau 2.  相似文献   

13.
A strain of Salmonella typhimurium, SO1007, which carries the amber mutation trpD28 plus the plasmid pKM101 was reverted very efficiently by two mutagens with different mutagenic specificities and modes of action: mitomycin C (MC) and N-methyl-N′-nitro-N-nitrosoguanidine (NG). By selecting revertants on minimal agar supplemented with anthranilic acid (AA), two distinct phenotypic classes of TrpD28 revertants can be recovered: prototrophs (MM+) and anthranilate utilizers (AA+). Since each phenotypic class is known to be caused by a variety of mutational events, reversion of trpD28 on minimal-anthranilate medium may be useful for detecting mutagenic agents regardless of the types of mutations they may cause. Thus, strains like SO 1007 may be useful as ‘universal’ detectors of mutagenic compounds. In the course of these experiments we also observed that pKM101 does not protect but, on the contrary, sensitizes the host bacteria slightly to the toxic effects of MC.  相似文献   

14.
Genetic variation at immunoglobulin (Ig) gene variable regions in B-cells is created through a multi-step process involving deamination of cytosine bases by activation-induced cytidine deaminase (AID) and their subsequent mutagenic repair. To protect the genome from dangerous, potentially oncogenic effects of off-target mutations, both AID activity and mutagenic repair are targeted specifically to the Ig genes. However, the mechanisms of targeting are unknown and recent data have highlighted the role of regulating mutagenic repair to limit the accumulation of somatic mutations resulting from the more widely distributed AID-induced lesions to the Ig genes. Here we investigated the role of the DNA damage sensor poly-(ADPribose)-polymerase-1 (PARP-1) in the repair of AID-induced DNA lesions. We show through sequencing of the diversifying Ig genes in PARP-1−/− DT40 B-cells that PARP-1 deficiency results in a marked reduction in gene conversion events and enhanced high-fidelity repair of AID-induced lesions at both Ig heavy and light chains. To further characterize the role of PARP-1 in the mutagenic repair of AID-induced lesions, we performed functional analyses comparing the role of engineered PARP-1 variants in high-fidelity repair of DNA damage induced by methyl methane sulfonate (MMS) and the mutagenic repair of lesions at the Ig genes induced by AID. This revealed a requirement for the previously uncharacterized BRCT domain of PARP-1 to reconstitute both gene conversion and a normal rate of somatic mutation at Ig genes, while being dispensable for the high-fidelity base excision repair. From these data we conclude that the BRCT domain of PARP-1 is required to initiate a significant proportion of the mutagenic repair specific to diversifying antibody genes. This role is distinct from the known roles of PARP-1 in high-fidelity DNA repair, suggesting that the PARP-1 BRCT domain has a specialized role in assembling mutagenic DNA repair complexes involved in antibody diversification.  相似文献   

15.
The generation of expression curves and the evaluation of mutagenic responses of mammalian cells using standard mutagenesis assays can be inaccurate because mutant and wild-type cells are usually mixed during the expression phase. If some mutant progenitors or mutants grow more slowly than the wild-type cells during the expression period, there will be a decrease in the mutant to wild-type ratio with time and the mutant fraction will not accurately represent the number of mutational events that occurred. The mutant fraction may also inaccurately assess the number of mutations if these mutations are expressed over a number of generations during the time before selection. We previously showed that recovery of L5178Y mouse cell mutants is not complete when mutations are allowed to express in suspension because slowly growing mutants and/or mutant progenitors are diluted out during this time (Rudd et al., 1990). In order to more accurately quantitate the mutagenic response of the cells, we developed an in situ procedure which segregates and immobilizes cells during expression. Because of this immobilization, slowly growing mutant progenitors and mutants expressed at different times will have an equal probability of being scored as mutants. Thus, one mutation leads to one mutant colony and the measurement of the mutagenic response of the cells to the chemical accurately reflects the mutational events that occurred. We plated L5178Y tk+/− mouse cells in semisolid medium immediately after treatment. As the cells grew and formed microcolonies, the selective agent TFT was added as an overlay at specified times, permitting only TFTr cells to survive. In this procedure, each mutation was captured as an individual colony; consequently, the measured mutation fraction accurately reflected the mutational events that occurred at the selected locus. In addition, the induced mutant colonies arising in the agar are the result of independent mutational events. We previously described the in situ protocol for L5178Y cells and showed that the spontaneous mutation rate measured was 50-fold greater than when the cells expressed the phenotype in suspension (Rudd et al., 1990). From this we concluded that the slow growth phenotype was expressed before TFT resistance. In the present paper, we evaluate the effect of chemical treatment on the mutation fraction as a function of the time to TFT addition. Using the in situ protocol, we generated expression curves for three nucleotide analogs, 5-azacytidine, TFT and AraC. The numbers of TFTr colonies produced at various times after treatment indicated that chemically-treated cultures had higher mutation fractions than the solvent controls. The maximal differential increase in mutation rate occured between 30 and 60 h for 5-azacytidine and between 20 and 40 h for TFT and AraC. Our results document the feasibility of quantitating induced mutation fractions using the in situ protocol, confirm the mutagenicity of AraC and 5azacytidine and demonstrate the mutagenic activity of TFT at the tk locus. In addition to recovering mutants more accurately than the suspension protocol, the in situ protocol has the advantage of being experimentally less labor and time intensive. Therefore, we believe that this method should be considered for evaluation as an assay to measure the potential mutagenic effects of chemicals in mammalian cells in vitro.  相似文献   

16.
Two genes of Aspergillus nidulans are known to function in UV mutagenesis, but have been assigned to different epistasis groups: uvsC, which is also required for meiosis and mitotic recombination, and uvsI, which may have no other function. To clarify their role in error-prone repair and to investigate their interaction, uvsI and uvsC single and uvsI;uvsC double mutant strains were further tested for mutagen sensitivities and characterized for effects on mutation. Spontaneous and induced frequencies were compared in forward and reverse mutation assays. All results confirmed that uvsI and uvsC are members of different epistasis groups, and demonstrated that these uvs mutants have very different defects in UV mutagenesis. The uvsI strains showed wild-type frequencies in all forward mutation tests, but greatly reduced spontaneous and UV-induced reversion of some, but not other, point mutations. In contrast, uvsC had similar effects in all assay systems: namely pronounced mutator effects and greatly reduced UV mutagenesis. Interestingly, the uvsI;uvsC double mutant strains differed from both single mutants; they clearly showed synergism for all types of reversion tested: none were ever obtained spontaneously, nor after induction by UV or EMS (ethylmethane sulfonate). Based on these results, we conclude that uvsI is active in a mutation-specific, specialized error-prone repair process in Aspergillus. In contrast, uvsC, which is now known to show sequence homology to recA, has a basic function in mutagenic UV repair in addition to recombinational repair, similar to recA of Escherichia coli.  相似文献   

17.
The mutagenic properties of acridines on pneumococcus are described. All seven acridines tested were mutagenic at the amiA locus conferring a resistance to 10-5 m aminopterin. The effects of quinacrine were more specifically investigated. It was observed that: mutants can be obtained only by treatment of exponentially growing cells; a sharp maximum mutagenic effect occurs at a concentration slightly lower than the bacteriostatic value; and the amount of quinacrine required to yield the maximum mutagenic effect decreases with the pH of the medium. Moreover, the number of mutants detected after quinacrine treatment varies from locus to locus. The majority of quinacrine-induced mutants are readily reverted by quinacrine, but not by nitrosoguanidine treatment. This suggests that in pneumococcus quinacrine induces mainly frameshift mutations. A further study of the revertants obtained by quinacrine treatment of quinacrine-induced mutants strengthens this interpretation: most of the revertants result from a mutation at the same site; some partial revertants exhibiting an intermediate resistance to aminopterin were found to contain two very closely linked mutated sites, each mutation conferring the maximum level of resistance to aminopterin. Thus, the majority of quinacrine-induced mutants at the amiA locus of pneumococcus consists of frameshift mutations. Nearly all of the isolated mutants induced by quinacrine as well as other acridines belong to the low efficiency class of transformation. It was concluded that the mismatch resulting from the pairing between the wild type and the frameshift-containing sequence is recognized by the excision-repair system involved in the discrimination function in a way similar to that in which it recognizes mismatched base pairs between a transition mutation and the wild-type sequence.  相似文献   

18.
The volatile composition of white Agudelo, Blanco lexitimo, Godello and red Serradelo cultivars (NW Spain) harvested at two different stages of ripening have been evaluated. C6-compounds, alcohols, volatile fatty acids, monoterpenes, C13-norisoprenoids, volatile phenols and carbonyl compounds were identified and quantified in free and glycosidically bound forms by gas chromatography–mass spectrometry (GC–MS). The total volatile concentration showed a significant increase between the two ripening stages studied for all cultivars. The free volatile composition increased during maturity for Godello and Serradelo cultivars; however the glycosidically bound concentration increases for all cultivars with exception of B. lexitimo. Free C6-compounds ((E)-2-hexanal, 1-hexanol and (E)-2-hexen-1-ol) and bound alcohols (benzyl alcohol and 2-phenylethanol) showed the highest concentrations of volatile compounds for all grape cultivars in the two dates studied. Godello cultivar showed the highest change of volatile concentration between two ripening dates because of the high value of free C6-compounds. B. lexitimo was the most terpene-rich cultivar at the last ripening stage due to linalool; however C13-norisoprenoids in free form were detected in low concentrations for all cultivars but not in Godello and B. lexitimo cultivars at the last ripening stage. Free hexanoic acid increased during ripening in all cultivars. The evolution of volatiles during ripening of grape juice from the cultivars studied was not proportional to the changes in sugar content, which shows that the technological and aromatic maturities did not occur at the same time in these cultivars. The results also showed the cultivar * ripening date interaction for all, free and bound, groups of compounds.  相似文献   

19.
Gene flow from recombinant-DNA-modified (GMO) trees is a major barrier to their public acceptance and regulatory approval. Because many intensively grown trees are vegetatively propagated, complete sexual sterility could be a powerful means to mitigate or prevent gene flow. We tested four pairs of zinc-finger nucleases (ZFNs) as mutagenic agents against the LEAFY and AGAMOUS orthologs in poplar that are expected to be required for sexual fertility. To reduce the potential for pleiotropic effects from mutagenesis, each of the pairs was functionally linked to a heat shock promoter to provide inducible ZFN expression. Using Agrobacterium tumefaciens, we transformed more than 21,000 total explants compromised of both male and female hybrid poplar. The rate of transformation for the ZFN constructs (2 %) was generally reduced compared to the transgenic control (8 %). We produced 391 ZFN transgenic shoots of which only two developed into plants with mutations in a target gene; both were 7-bp deletions in one allele of the PtAG2 locus. No mutations were observed in the PtAG1 or PtLFY loci. Our results indicate a mutation rate of zero to 0.3 % per explant per allele, among the lowest reported for ZFN mutagenesis in plants. The combined effects of low recovery of transgenic plants, a modest mutation frequency, and much higher reported rates of directed mutation for other gene editing methods suggest that the efficient use of ZFNs in poplar requires further technical improvements.  相似文献   

20.
In this study, we present an efficient phosphorylation-free and ligase-free PCR-based multiple site-directed mutagenesis that allows simultaneous mutations up to six distal sites. This method could be extended to any plasmid DNA that is isolated from dam+Escherichia coli strains, and the results showed that the simultaneously mutagenic efficiencies of quadruple mutation and sextuple mutation were up to 80% and 40%, respectively.  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