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.
Oenococcus oeni is an alcohol-tolerant, acidophilic lactic acid bacterium primarily responsible for malolactic fermentation in wine. A recent comparative genomic analysis of O. oeni PSU-1 with other sequenced lactic acid bacteria indicates that PSU-1 lacks the mismatch repair (MMR) genes mutS and mutL. Consistent with the lack of MMR, mutation rates for O. oeni PSU-1 and a second oenococcal species, O. kitaharae, were higher than those observed for neighboring taxa, Pediococcus pentosaceus and Leuconostoc mesenteroides. Sequence analysis of the rpoB mutations in rifampin-resistant strains from both oenococcal species revealed a high percentage of transition mutations, a result indicative of the lack of MMR. An analysis of common alleles in the two sequenced O. oeni strains, PSU-1 and BAA-1163, also revealed a significantly higher level of transition substitutions than were observed in other Lactobacillales species. These results suggest that the genus Oenococcus is hypermutable due to the loss of mutS and mutL, which occurred with the divergence away from the neighboring Leuconostoc branch. The hypermutable status of the genus Oenococcus explains the observed high level of allelic polymorphism among known O. oeni isolates and likely contributed to the unique adaptation of this genus to acidic and alcoholic environments.  相似文献   

2.
Summary Previous studies have shown that the mutT, mutH and mutL mutators of Escherichia coli have a marked advantage in competition growth with otherwise coisogenic wild-type strains. As shown in this paper the same is true for the mutS mismatch mutator. In three experiments mutS could outgrow the wild-type and had higher fitness values.  相似文献   

3.
Some strains of the soil bacterium Pseudomonas putida have become in recent years platforms of choice for hosting biotransformations of industrial interest. Despite availability of many genetic tools for this microorganism, genomic editing of the cell factory P. putida EM42 (a derivative of reference strain KT2440) is still a time‐consuming endeavor. In this work we have investigated the in vivo activity of the Ssr protein encoded by the open reading frame T1E_1405 from Pseudomonas putida DOT‐T1E, a plausible functional homologue of the β protein of the Red recombination system of λ phage of Escherichia coli. A test based on the phenotypes of pyrF mutants of P. putida (the yeast's URA3 ortholog) was developed for quantifying the ability of Ssr to promote invasion of the genomic DNA replication fork by synthetic oligonucleotides. The efficiency of the process was measured by monitoring the inheritance of the changes entered into pyrF by oligonucleotides bearing mutated sequences. Ssr fostered short and long genomic deletions/insertions at considerable frequencies as well as single‐base swaps not affected by mismatch repair. These results not only demonstrate the feasibility of recombineering in P. putida, but they also enable a suite of multiplexed genomic manipulations in this biotechnologically important bacterium.  相似文献   

4.
Defects in DNA mismatch repair (MMR) occur frequently in natural populations of pathogenic and commensal bacteria, resulting in a mutator phenotype. We identified a unique genetic element in Streptococcus pyogenes strain SF370 that controls MMR via a dynamic process of prophage excision and reintegration in response to growth. In S. pyogenes, mutS and mutL are organized on a polycistronic mRNA under control of a common promoter. Prophage SF370.4 is integrated between the two genes, blocking expression of the downstream gene (mutL) and resulting in a mutator phenotype. However, in rapidly growing cells the prophage excises and replicates as an episome, allowing mutL to be expressed. Excision of prophage SF370.4 and expression of MutL mRNA occur simultaneously during early logarithmic growth when cell densities are low; this brief window of MutL gene expression ends as the cell density increases. However, detectable amounts of MutL protein remain in the cell until the onset of stationary phase. Thus, MMR in S. pyogenes SF370 is functional in exponentially growing cells but defective when resources are limiting. The presence of a prophage integrated into the 5′ end of mutL correlates with a mutator phenotype (10−7 to 10−8 mutation/generation, an approximately a 100-fold increase in the rate of spontaneous mutation compared with prophage-free strains [10−9 to 10−10 mutation/generation]). Such genetic elements may be common in S. pyogenes since 6 of 13 completed genomes have related prophages, and a survey of 100 strains found that about 20% of them are positive for phages occupying the SF370.4 attP site. The dynamic control of a major DNA repair system by a bacteriophage is a novel method for achieving the mutator phenotype and may allow the organism to respond rapidly to a changing environment while minimizing the risks associated with long-term hypermutability.  相似文献   

5.
DNA base sequence changes induced by bromouracil mutagenesis of lambda phage   总被引:14,自引:0,他引:14  
The base sequence changes induced by bromouracil mutagenesis in the cI gene of phage lambda have been determined by direct sequence analysis. Phage DNA mutagenized during prophage replication or during phage lytic growth showed predominantly A · T → G · C transitions. The frequency of this mutation was strongly sequence-dependent: 5′ A-C-G-C 3′ > A-C(A.C or T) > A(A.G or T). The difference in mutability of bases in the gene is not the result of specificity in mutL-dependent mismatch repair, since phage grown in mutL host cells showed the same distribution of bromouracil mutations. The observations made in phage mutagenized with bromouracil in the prophage state should be representative of bromouracil mutagenesis in the Escherichia coli chromosome.  相似文献   

6.
Mutator mutations in the mutS gene induced by the insertion of phage Mu or the transposable resistance elements Tn5 or Tn10 and those in the mutL gene induced by Tn5 and T10 gave mutagenic activities similar to that of the previously described mutS3 and mutL25 mutations. Various combinations of mutS::Tn5,mutL::Tn5, uvrE156, and the deletion mutation δmutH2 did not produce an additive effect. This supports the idea that the products of these genes function in the same pathway of error correction during DNA synthesis.  相似文献   

7.
Pseudomonas putida has emerged as a promising host for the production of chemicals and materials thanks to its metabolic versatility and cellular robustness. In particular, P. putida KT2440 has been officially classified as a generally recognized as safe (GRAS) strain, which makes it suitable for the production of compounds that humans directly consume, including secondary metabolites of high importance. Although various tools and strategies have been developed to facilitate metabolic engineering of P. putida, modification of large genes/clusters essential for heterologous expression of natural products with large biosynthetic gene clusters (BGCs) has not been straightforward. Recently, we reported a RecET-based markerless recombineering system for engineering P. putida and demonstrated deletion of multiple regions as large as 101.7 kb throughout the chromosome by single rounds of recombineering. In addition, development of a donor plasmid system allowed successful markerless integration of heterologous BGCs to P. putida chromosome using the recombineering system with examples of – but not limited to – integrating multiple heterologous BGCs as large as 7.4 kb to the chromosome of P. putida KT2440. In response to the increasing interest in our markerless recombineering system, here we provide detailed protocols for markerless gene knockout and integration for the genome engineering of P. putida and related species of high industrial importance.  相似文献   

8.
The mismatch repair system is involved in the maintenance of genomic integrity by editing DNA replication and recombination. However, although most mutations are neutral or deleterious, a mutator phenotype due to an inefficient mismatch repair may generate advantageous variants and may therefore be selected for. We review the evidence for inefficient mismatch repair due either to genetic defects in mismatch repair genes or to physiological conditions. Among natural isolates ofEscherichia coli andSalmonella enterica, about 1% are mutator bacteria, mostly deficient in mismatch repair (most of them defective in themutS gene). Characterization of mutators derived from laboratory strains led also to the isolation of mismatch repair mutants in which the most frequently found defects are inmutL andmutS. The correlation of the size of the antimutator genes with the frequency of their defective alleles amongE. coli andSalmonella strains reveals thatmutU mutants are underrepresented. Analysis of the progeny of a defined M13 phage heteroduplex DNA transfected intoE. coli cells shows that mismatch repair efficiency progressively decreases from the end of the exponential growth in K-12 and is variable among natural isolates. Implications of this defective mismatch repair activity for evolution and tumorigenesis will be discussed.  相似文献   

9.
Genetic stability depends in part on an efficient DNA lesion recognition and correction by the DNA mismatch repair (MMR) system. In eukaryotes, MMR is initiated by the binding of heterodimeric MutS homologue (MSH) complexes, MSH2–MSH6 and MSH2–MSH3, which recognize and bind mismatches and unpaired nucleotides. Plants encode another mismatch recognition protein, named MSH7. MSH7 forms a heterodimer with MSH2 and the protein complex is designated MutSγ. We here report the effect the expression of Arabidopsis MSH2 and MSH7 alone or in combination exert on the genomic stability of Saccharomyces cerevisiae. AtMSH2 and AtMutSγ proteins failed to complement the hypermutator phenotype of an msh2 deficient strain. However, overexpressing AtMutSγ in MMR proficient strains generated a 4-fold increase in CAN1 forward mutation rate, when compared to wild-type strains. Canr mutation spectrum analysis of AtMutSγ overproducing strains revealed a substantial increase in the frequency of base substitution mutations, including an increased accumulation of base pair changes from G:C to A:T and T:A to C:G, G:C or A:T. Taken together, these results suggest that AtMutSγ affects yeast genomic stability by recognizing specific mismatches and preventing correction by yeast MutSα and MutSβ, with subsequent inability to interact with yeast downstream proteins needed to complete MMR.  相似文献   

10.
Genetic stability relies in part on the efficiency with which post-replicative mismatch repair (MMR) detects and corrects DNA replication errors. In Escherichia coli, endogenous transition mispairs and insertion/deletion (ID) heterologies are corrected with similar efficiencies – but much more efficiently than transversion mispairs – as revealed by mutation rate increases in MMR mutants. To assess the relative efficiencies with which these mismatches are corrected in the yeast Saccharomyces cerevisiae, we examined repair of defined mismatches on heteroduplex plasmids and compared the spectra for >1000 spontaneous SUP4-o mutations arising in isogenic wild-type or MMR-deficient (pms1, mlh1, msh2) strains. Heteroduplexes containing G/T mispairs or ID heterologies were corrected more efficiently than those containing transversion mismatches. However, the rates of single base-pair insertion/deletion were increased much more (82-fold or 34-fold, respectively) on average than the rate of base pair substitutions (4.4-fold), with the rates for total transitions and transversions increasing to similar extents. Thus, the relative efficiencies with which mismatches formed during DNA replication are repaired appear to differ in prokaryotic and eukaryotic cells. In addition, our results indicate that in yeast, and probably other eukaryotes, these efficiencies may not mirror those obtained from an analysis of heteroduplex correction. Received: 15 November 1998 / Accepted: 4 February 1999  相似文献   

11.
Implementation of single-stranded DNA (ssDNA) recombineering in Pseudomonas putida has widened the range of genetic manipulations applicable to this biotechnologically relevant bacterium. Yet, the relatively low efficiency of the technology hampers identification of mutated clones lacking conspicuous phenotypes. Fortunately, the use of CRISPR/Cas9 as a device for counterselection of wild-type sequences helps to overcome this limitation. Merging ssDNA recombineering with CRISPR/Cas9 thus enables a suite of genomic edits with a straightforward approach: a CRISPR plasmid provides the spacer DNA sequence that directs the Cas9 nuclease ribonucleoprotein complex to cleave the genome at the wild-type sequences that have not undergone the change entered by the mutagenic ssDNA oligonucleotide(s). This protocol describes a complete workflow of the method optimized for P. putida, although it could in principle be applicable to many other pseudomonads. As an example, we show the deletion of the edd gene that encodes one key enzyme that operates the EDEMP cycle for glucose metabolism in P. putida EM42. By combining two incompatible CRISPR plasmids with different antibiotic selection markers, we show that the procedure can be cycled to implement consecutive deletions in the same strain, e.g. deletion of the pyrF gene following that of the edd mutant. This approach adds to the wealth of genetic technologies available for P. putida and strengthens its status as a chassis of choice for a suite of biotechnological applications.  相似文献   

12.
We have characterized the mismatch repair system (MMR) of the highly radiation-resistant type strain of Deinococcus radiodurans, ATCC 13939. We show that the MMR system is functional in this organism, where it participates in ensuring the fidelity of DNA replication and recombination. The system relies on the activity of two key proteins, MutS1 and MutL, which constitute a conserved core involved in mismatch recognition. Inactivation of MutS1 or MutL resulted in a seven-fold increase in the frequency of spontaneous RifR mutagenesis and a ten-fold increase in the efficiency of integration of a donor point-mutation marker during bacterial transformation. Inactivation of the mismatch repair-associated UvrD helicase increased the level of spontaneous mutagenesis, but had no effect on marker integration—suggesting that binding of MutS1 and MutL proteins to a mismatched heteroduplex suffices to inhibit recombination between non identical (homeologous) DNAs. In contrast, inactivation of MutS2, encoded by the second mutS -related gene present in D. radiodurans, had no effect on mutagenesis or recombination. Cells devoid of MutS1 or MutL proteins were as resistant to -rays, mitomycin C and UV-irradiation as wild-type bacteria, suggesting that the mismatch repair system is not essential for the reconstitution of a functional genome after DNA damage.Electronic Supplementary Material Supplementary material is available in the online version of this article at Communicated by G. Baldacci  相似文献   

13.
The activities of the Vsr and MutH endonucleases of Escherichia coli are stimulated by MutL. The interaction of MutL with each enzyme is enhanced in vivo by 2-aminopurine treatment and by inactivation of the mutY gene. We hypothesize that MutL recruits the endonucleases to sites of DNA damage.The Escherichia coli Dcm protein methylates the second C of CCWGG sites (W = A or T). Deamination of 5-methylcytosine converts CG base pairs to T/G mismatches, causing CCWGG-to-CTWGG transition mutations. Very-short-patch (VSP) repair minimizes these mutations (2). Repair is initiated by a sequence- and mismatch-specific endonuclease, Vsr, which cleaves the DNA 5′ of the T. DNA polymerase I removes the T along with a few 3′ nucleotides and resynthesizes the missing bases, restoring the CG base pair. Vsr is both necessary and sufficient for initiating VSP repair. However, two other proteins, MutS and MutL, enhance VSP repair of deamination damage (1).MutS and MutL are best known for their roles in postreplication mismatch repair (MMR) (9, 11). MutL couples mismatch recognition by MutS to the activation of MutH, an endonuclease that cleaves the unmethylated strand of GATC sequences that are transiently hemimethylated following DNA replication. The nicked strand, containing the erroneous base, is removed by the UvrD helicase and one of several exonucleases to beyond the mismatch and then resynthesized by DNA polymerase III.MutL stimulates the endonuclease activities of both Vsr and MutH in vitro (8, 17). The requirements for stimulation are the same: a mismatch, MutS, and ATP hydrolysis by MutL (8, 8a). Cross-linking studies showed that MutH and Vsr interact with the same region in the N-terminal domain of MutL (Heinze et al., submitted). Competition of Vsr with MutH for access to MutL explains the ability of Vsr to inactivate MMR in vivo when overexpressed (6, 13). Thus, the interactions of the two repair endonucleases with MutL are structurally and functionally very similar.In contrast to MMR, where the cleavage site for MutH may be several kilobases away from the mismatch, VSP repair requires that mismatch recognition and endonucleolytic cleavage occur at the same C(T/G)WGG site. How MutS and MutL stimulate VSP repair if MutS and Vsr compete for the same mismatch remains unknown (2, 12). We hypothesized that MutS binds the mismatch first and that a MutS-MutL complex then recruits Vsr. If so, then the MMR proteins would initially mask the mismatch, making the interaction of Vsr with MutL independent of lesion identity.To test this hypothesis, we studied the interaction of MutL with Vsr and with MutH in response to two types of mismatch by using a bacterial two-hybrid assay (10). This assay detects all known interactions among the Mut proteins: homodimerization of MutS and MutL, interaction of MutL with MutS and with MutH, and interaction of Vsr with the N-terminal domain of MutL (15). We found no false positives or false negatives. Furthermore, since the assay relies on reconstitution of a soluble protein (adenylate cyclase), the DNA repair proteins are free to interact with the DNA (Fig. (Fig.11).Open in a separate windowFIG. 1.Known interactions among repair proteins as detected by the bacterial two-hybrid assay. The T18 and T25 subunits of CyaA are fused to any two repair proteins (illustrated here by MutL and Vsr), allowing measurement of all pairwise interactions as units of β-galactosidase (β-gal). T25 fusions are repair proficient. CRP, cyclic AMP (cAMP) receptor protein; P, lac operon promoter; RNAP, RNA polymerase.2-Aminopurine (2AP) mispairs with C during DNA replication, causing transition and frameshift mutations (5). The transitions are due primarily to the mismatch itself; the frameshifts are due to saturation of MMR, which leaves slipped-strand intermediates caused by DNA replication errors unrepaired (19). MutS and MutL bind to 2AP/C lesions (22), although the lesions may not be subject to MMR (19). As shown in Fig. Fig.2,2, treatment with 2AP causes a dose-dependent increase in the interaction of MutL with both Vsr and MutH; dimerization of MutL and interaction of MutL with MutS are somewhat increased.Open in a separate windowFIG. 2.Effect of 2AP treatment on protein-protein interactions in the bacterial two-hybrid assay. Results in units of β-galactosidase ± standard errors of the means (n = 9) are shown for BTH101(F galE15 ga1K16 rpsL1 hsdR2 mcrA1 mcrB1 cyaA-99) cells treated with 2AP as described previously (5, 19). Cells were cotransformed with pT18 and pT25 vectors (light gray bars), pT18-mutS and pT25-mutL (white bars), pT18-vsr and pT25-mutL (gray bars), pT18-mutH and pT25-mutL (black bars), or pT18-mutL and pT25-mutL (mottled bars). (NB: The dose-response curve for the pT18-mutS pT25-mutS transformants is similar to that of the pT18-mutL pT25-mutL transformants; it has been omitted for graphical clarity since the MutS-MutS interaction gives very high units of β-galactosidase activity [15]).The MutY adenine glycosylase removes A''s which have mispaired with oxidized guanine (8-oxoG) during DNA replication. Cells with a deletion of mutY have an elevated frequency of CG-to-AT transversion mutations (18); these are reduced by excess MutS, suggesting that 8-oxoG/A mismatches are also subject to MMR (23). As shown in Fig. Fig.3,3, the interactions between Vsr and MutL and between MutH and MutL increase in a mutY cell (stippled bars). Other interactions, such as MutS dimerization, are unaffected (not shown).Open in a separate windowFIG. 3.Effects of mutY and mutT deletions on protein-protein interactions in the bacterial two-hybrid assay. Results are in units of β-galactosidase, relative to the level in the wild type, in mutT (solid) and mutY (stippled) derivatives of BTH101 cotransformed with pT18 and pT25 vectors, pT18-mutH and pT25-mutL, pT18-vsr and pT25-mutL, or pT18-mutS and pT25-mutS (n = 3).8-OxoG/A mismatches also arise by incorporation of oxidized dGTP opposite A during DNA replication. The MutT nuclease minimizes this by removing oxidized dGTP from the nucleotide pool. The high frequency of AT-to-CG mutations in mutT strains is unaffected by the status of the MMR system (7, 21, 23), possibly because these 8-oxoG/A mispairs are in a conformation that MutS does not recognize. As shown in Fig. Fig.3,3, neither the interaction between MutL and Vsr nor that between MutL and MutH is elevated in a mutT strain (solid bars).These data show that mismatches which attract MutS and MutL increase the interaction of MutL with MutH in vivo. Although these mismatches are not subject to VSP repair, they also increase the interaction between MutL and Vsr. The simplest interpretation is that a MutS-MutL complex recruits MutH and Vsr to the DNA independent of the identity of the mismatch. MutS and MutL could then clear the mismatch, delivering the (activated) endonuclease to its specific target site, no matter how far away it is.Interaction of MutL with MutH, leading to MMR, is probably the default option. However, the MutS-MutL complex may recruit other repair proteins, such as Vsr or UvrB (20), to lesions that are poorly processed by MMR. The T/G mismatch in hemimethylated CTWGG sequences may be one such site. Vsr is expressed at very low levels in growing cells (14), so this recruitment would enhance VSP repair. However, recruitment of Vsr to other lesions would reduce VSP repair. For example, recruitment of Vsr by MutL to 2AP/C lesions (Fig. (Fig.2)2) could explain why CCWGG sites are hotspots for 2AP-induced mutations (4, 19).We have argued that Vsr is kept at low levels while DNA is replicating to avoid interference with MMR (14). However, if, as we suggest here, MutS and MutL are needed to recruit scarce Vsr to its target sequence, this argument loses its merit. It seems more likely that Vsr levels are kept low to avoid CTWGG-to-CCWGG mutations; Vsr creates these mutations by converting T/G mismatches formed at CTAGG sites by errors in DNA replication to CG (3, 6, 16). Vsr levels rise in nongrowing cells (14), when mutagenesis is no longer a risk. Under these circumstances, it is likely that MutS and MutL are no longer required for efficient VSP repair.  相似文献   

14.
The most important system for correcting replication errors that survive the built in editing system of DNA polymerase is the mismatch repair (MMR) system. We have identified a novel mutator strain yycJ in Bacillus anthracis. Mutations in the yycJ gene result in a spontaneous mutator phenotype with a mutational frequency and specificity comparable to that of MMR-deficient strains such as those with mutations in mutL or mutS. YycJ was annotated as a metallo-β-lactamase (MβL) super family member with unknown activity. In this study we carried out a biochemical characterization of YycJ and demonstrated that a recombinant YycJ protein possesses a 5′–3′ exonuclease activity at the 5′ termini and at nicks of double-stranded DNA. This activity requires a divalent metal cofactor Mn2+ and is stimulated by 5′-phosphate ends of duplex DNA. The mutagenesis of conserved amino acid residues revealed that in addition to the five MβL family conserved motifs, YycJ appears to have its specific motifs that can be used to distinguish YycJ from other closely related MβL family members. A phylogenetic survey showed that putative YycJ homologs are present in several bacterial phyla as well as in members of the Methanomicrobiales and Thermoplasmales from Archaea. We propose that YycJ represents a new group of MβL fold exonucleases, which is likely to act in the recognition of MMR entry point and subsequent removal of the mismatched base in certain MutH-less bacterial species.  相似文献   

15.
Escherichia coli MutS is a highly conserved mismatch repair (MMR) protein that plays a key role in recognizing DNA mismatches and the early steps of MMR. Previous studies revealed an interaction between MutS and the replicative protein β clamp, but it remains unclear whether the interaction functions during the process of MMR. In order to provide insight into the significance of this interaction, Far Western, Surface plasmon resonance and cell survival/mutagenesis assays were used to determine its possible influences on the in vitro and in vivo properties of MutS. The results show that a quintuple mutation of MutS residues 812–816 (MutSβC), or single alanine substitution mutation of MutS residues M813 or L815 completely blocks binding of MutS to β clamp. Wild type β clamp interferes with DNA binding by MutS. When treated with the base analog 2-aminopurine, MutSβC confers more mutations and less cellular growth rate in the mutS-deficient strain than the wild type MutS. These data indicate that the MutS-β interaction has functional consequences during MMR in E. coli.  相似文献   

16.
Colonies of Bacillus anthracis Sterne allow the growth of papillation after 6 days of incubation at 30°C on Luria–Bertani medium. The papillae are due to mutations that allow the cells to overcome the barriers to continued growth. Cells isolated from papillae display two distinct gross phenotypes (group A and group B). We determined that group A mutants have mutations in the nprR gene including frameshifts, deletions, duplications and base substitutions. We used papillation as a tool for finding new mutators as the mutators generate elevated levels of papillation. We discovered that disruption of yycJ or recJ leads to a spontaneous mutator phenotype. We defined the nprR/papillation system as a new mutational analysis system for B. anthracis. The mutational specificity of the new mutator yycJ is similar to that of mismatch repair‐deficient strains (MMR) such as those with mutations in mutL or mutS. Deficiency in recJ results in a unique specificity, generating only tandem duplications.  相似文献   

17.
Mismatch repair (MMR) corrects replication errors such as mismatched bases and loops in DNA. The evolutionarily conserved dimeric MMR protein MutS recognizes mismatches by stacking a phenylalanine of one subunit against one base of the mismatched pair. In all crystal structures of G:T mismatch-bound MutS, phenylalanine is stacked against thymine. To explore whether these structures reflect directional mismatch recognition by MutS, we monitored the orientation of Escherichia coli MutS binding to mismatches by FRET and anisotropy with steady state, pre-steady state and single-molecule multiparameter fluorescence measurements in a solution. The results confirm that specifically bound MutS bends DNA at the mismatch. We found additional MutS-mismatch complexes with distinct conformations that may have functional relevance in MMR. The analysis of individual binding events reveal significant bias in MutS orientation on asymmetric mismatches (G:T versus T:G, A:C versus C:A), but not on symmetric mismatches (G:G). When MutS is blocked from binding a mismatch in the preferred orientation by positioning asymmetric mismatches near the ends of linear DNA substrates, its ability to authorize subsequent steps of MMR, such as MutH endonuclease activation, is almost abolished. These findings shed light on prerequisites for MutS interactions with other MMR proteins for repairing the appropriate DNA strand.  相似文献   

18.
 This study examined which culture fraction of the plant-growth-promoting bacterium Pseudomonas putida (Trevisan) Migula has an effect on growth and mycorrhiza formation of maize (Zea mays L.). Shoot dry weight and total leaf area of plants did not increase after inoculation with Glomus fistulosum but were significantly higher than the controls when the plants were dualinoculated with G. fistulosum and living cells of P. putida. Mycorrhizal infection of the roots was significantly higher when plants were inoculated with G. fistulosum together with living cells of P. putida or with G. fistulosum and dialysed cell extracts of P. putida than with G. fistulosum alone. Development of arbuscular mycorrhizal (AM) extraradical hyphae and the proportion of extraradical hyphae showing NADH diaphorase activity were significantly enhanced by inoculation of plants with living cells of P. putida or dialysed cell extracts of P. putida. No stimulation of extraradical hyphae proliferation from in vitro incubated mycorrhizal root segments was observed after application of culture fractions of P. putida. However, the percentage contamination of the root segments by extraneous filamentous fungi significantly decreased in the presence of livingcells of P. putida. Accepted: 12 January 1996  相似文献   

19.
The cytosine methyltransferases (MTases) M. HhaIand M. HpaII bind substrates in which the target cytosine is replaced by uracil or thymine, i.e. DNA containing a U:G or a T:G mismatch. We have extended this observation to the EcoRII MTase (M. EcoRII) and determined the apparent Kd for binding. Using a genetic assay we have also tested the possibility that MTase binding to U:G mismatches may interfere with repair of the mismatches and promote C:G to T:A mutations. We have compared two mutants of M. EcoRII that are defective for catalysis by the wild-type enzyme for their ability to bind DNA containing U:G or T:G mismatches and for their ability to promote C to T mutations. We find that although all three proteins are able to bind DNAs with mismatches, only the wild-type enzyme promotes C:G to T:A mutations in vivo. Therefore, the ability of M. EcoRII to bind U:G mismatched duplexes is not sufficient for their mutagenic action in cells. Received: 14 November 1996 / Accepted: 17 February 1997  相似文献   

20.
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