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.
We have previously shown that the assembly of RecA protein onto single-stranded DNA (ssDNA) facilitated by SSB protein occurs in three steps: (1) rapid binding of SSB protein to the ssDNA; (2) nucleation of RecA protein onto this template; and (3) co-operative polymerization of additional RecA protein to yield presynaptic filaments. Here, electron microscopy has been used to further explore the parameters of this assembly process. The optimal extent of presynaptic filament formation required at least one RecA protein monomer per three nucleotides, high concentrations of ATP (greater than 3 mM in the presence of 12 mM-Mg2+), and relatively low concentrations of SSB protein (1 monomer per 18 nucleotides). Assembly was depressed threefold when SSB protein was added to one monomer per nine nucleotides. These effects appeared to be exerted at the nucleation step. Following nucleation, RecA protein assembled onto ssDNA at net rates that varied from 250 to 900 RecA protein monomers per minute, with the rate inversely related to the concentration of SSB protein. Combined sucrose sedimentation and electron microscope analysis established that SSB protein was displaced from the ssDNA during RecA protein assembly.  相似文献   

2.
The RecA protein from Escherichia coli promotes an ATP-dependent three-strand exchange reaction between a circular single-stranded DNA (ssDNA) and a homologous linear double-stranded (dsDNA). We have now found that under certain conditions, the RecA protein is also able to promote the three-strand exchange reaction using the structurally related nucleoside triphosphate, ITP, as the nucleotide cofactor. However, although both reactions are stimulated by single-stranded DNA-binding (SSB) protein, the ITP-dependent reaction differs from the ATP-dependent reaction in that it is observed only at low SSB protein concentrations, whereas the ATP-dependent reaction proceeds efficiently even at high SSB protein concentrations. Moreover, the circular ssDNA-dependent ITP hydrolysis activity of the RecA protein is strongly inhibited by SSB protein (suggesting that SSB protein displaces RecA protein from ssDNA when ITP is present), whereas the ATP hydrolysis activity is uninhibited even at high SSB protein concentrations (because RecA protein is resistant to displacement by SSB protein when ATP is present). These results suggest that SSB protein does not stimulate the ITP-dependent strand exchange reaction presynaptically (by facilitating the binding of RecA protein to the circular ssDNA substrate) but may act postsynaptically (by binding to the displaced strand that is generated when the circular ssDNA invades the linear dsDNA substrate). Interestingly, the mechanistic characteristics of the ITP-dependent strand exchange reaction of the E. coli RecA protein are similar to those of the ATP-dependent strand exchange reaction of the RecA protein from Streptococcus pneumoniae. These findings are discussed in terms of the relationship between the dynamic state of the RecA-ssDNA filament and the mechanism of the SSB protein-stimulated three-strand exchange reaction.  相似文献   

3.
We have characterized the double-stranded DNA (dsDNA) binding properties of RecA protein, using an assay based on changes in the fluorescence of 4',6-diamidino-2-phenylindole (DAPI)-dsDNA complexes. Here we use fluorescence, nitrocellulose filter-binding, and DNase I-sensitivity assays to demonstrate the binding of two duplex DNA molecules by the RecA protein filament. We previously established that the binding stoichiometry for the RecA protein-dsDNA complex is three base-pairs per RecA protein monomer, in the presence of ATP. In the presence of ATPgammaS, however, the binding stoichiometry depends on the MgCl2 concentration. The stoichiometry is 3 bp per monomer at low MgCl2 concentrations, but changes to 6 bp per monomer at higher MgCl2 concentrations, with the transition occurring at approximately 5 mM MgCl2. Above this MgCl2 concentration, the dsDNA within the RecA nucleoprotein complex becomes uncharacteristically sensitive to DNase I digestion. For these reasons we suggest that, at the elevated MgCl2 conditions, the RecA-dsDNA nucleoprotein filament can bind a second equivalent of dsDNA. These results demonstrate that RecA protein has the capacity to bind two dsDNA molecules, and they suggest that RecA or RecA-like proteins may effect homologous recognition between intact DNA duplexes.  相似文献   

4.
The middle and high molecular weight members of the neurofilament triplet, NF-M and NF-H, undergo extensive posttranslational polyphosphorylation, a process requiring 24 h or more for completion. We have investigated ways of perturbing this process in intact cells and have found that phosphorylation of newly synthesized NF-M in cultured chick sensory neurons is inhibited by Li+. [35S]Methionine pulse-chase experiments were carried out with pure neuronal cultures, and the phosphorylation of newly synthesized NF-M was monitored by following the accompanying change, with chase time, in apparent size and charge of the polypeptide. Addition of LiCl to the medium inhibited this mobility shift in a dose-dependent manner over concentrations between 2 and 25 mM. Incorporation of 32P into NF-M, as well as NF-H, was also inhibited, whereas incorporation into the low molecular weight neurofilament protein, beta-tubulin, and total protein was unaffected. Protein synthesis was not altered. Exposure to 25 mM LiCl for up to 72 h was not toxic, and the inhibition of NF-M phosphorylation was completely reversible. When 25 mM Li+ was added after NF-M had become partially phosphorylated, further progression was blocked, but there was no net dephosphorylation or degradation of NF-M. Additional experiments suggest that this action of Li+ is probably not due to effects on second messenger levels or to effects on tubulin metabolism and assembly state presented in our accompanying article, but rather to interference by Li+ itself, with the phosphorylation of NF-M and NF-H by specific neurofilament kinase(s).  相似文献   

5.
The ATP-dependent three-strand exchange activity of the Streptococcus pneumoniae RecA protein (RecA(Sp)), like that of the Escherichia coli RecA protein (RecA(Ec)), is strongly stimulated by the single-stranded DNA-binding protein (SSB) from either E. coli (SSB(Ec)) or S. pneumoniae (SSB(Sp)). The RecA(Sp) protein differs from the RecA(Ec) protein, however, in that its ssDNA-dependent ATP hydrolysis activity is completely inhibited by SSB(Ec) or SSB(Sp) protein, apparently because these proteins displace RecA(Sp) protein from ssDNA. These results indicate that in contrast to the mechanism that has been established for the RecA(Ec) protein, SSB protein does not stimulate the RecA(Sp) protein-promoted strand exchange reaction by facilitating the formation of a presynaptic complex between the RecA(Sp) protein and the ssDNA substrate. In addition to acting presynaptically, however, it has been proposed that SSB(Ec) protein also stimulates the RecA(Ec) protein strand exchange reaction postsynaptically, by binding to the displaced single strand that is generated when the ssDNA substrate invades the homologous linear dsDNA. In the RecA(Sp) protein-promoted reaction, the stimulatory effect of SSB protein may be due entirely to this postsynaptic mechanism. The competing displacement of RecA(Sp) protein from the ssDNA substrate by SSB protein, however, appears to limit the efficiency of the strand exchange reaction (especially at high SSB protein concentrations or when SSB protein is added to the ssDNA before RecA(Sp) protein) relative to that observed under the same conditions with the RecA(Ec) protein.  相似文献   

6.
Oxygen as attractant and repellent in bacterial chemotaxis.   总被引:12,自引:8,他引:4       下载免费PDF全文
J Shioi  C V Dang    B L Taylor 《Journal of bacteriology》1987,169(7):3118-3123
Studies of bacterial chemotaxis to oxygen (aerotaxis) over a broad range of oxygen concentrations showed that at high concentrations, oxygen was a repellent of Salmonella typhimurium, Escherichia coli, and some bacilli, whereas it is known that at lower concentrations (less than or equal to 0.25 mM dissolved oxygen), oxygen is an attractant. In a temporal assay of aerotaxis, S. typhimurium in medium equilibrated with air (0.25 mM dissolved oxygen) and then exposed to pure oxygen (1.2 mM) tumbled continuously for approximately 20 s. The oxygen concentration that elicited a half-maximal negative (repellent) response was 1.0 mM for both S. typhimurium and E. coli. The receptor for the negative chemoresponse to high concentrations of oxygen is apparently different from the receptor for the positive chemoresponse to low concentrations of oxygen, since the oxygen concentration that elicits a half-maximal positive (attractant) response in S. typhimurium and E. coli is reported to be 0.7 microM. Adaptation to high concentrations of oxygen, like adaptation to low concentrations of oxygen, was independent of methylation of a transducer protein. Only the response to low oxygen concentrations, however, was altered by interaction with the amidated Tsr transducer in cheB mutants.  相似文献   

7.
The effects of leucine, its metabolites, and the 2-oxo acids of valine and isoleucine on protein synthesis and degradation in incubated limb muscles of immature and adult rats were tested. Leucine stimulated protein synthesis but did not reduce proteolysis when leucine transamination was inhibited. 4-Methyl-2-oxopentanoate at concentrations as low as 0.25 mM inhibited protein degradation but did not change protein synthesis. The 2-oxo acids of valine and isoleucine did not change protein synthesis or degradation even at concentrations as high as 5 mM. 3-Methylvalerate, the irreversibly decarboxylated product of 4-methyl-2-oxopentanoate, decreased protein degradation at concentrations greater than or equal to 1 mM. This was not due to inhibition of 4-methyl-2-oxopentanoate catabolism, because 0.5 mM-3-methylvalerate did not suppress proteolysis, even though it inhibited leucine decarboxylation by 30%; higher concentrations of 3-methylvalerate decreased proteolysis progressively without inhibiting leucine decarboxylation further. During incubation with [1-14C]- and [U-14C]-leucine, it was found that products of leucine catabolism formed subsequent to the decarboxylation of 4-methyl-2-oxopentanoate accumulated intracellularly. This pattern was not seen during incubation with radiolabelled valine. Thus, the effect of leucine on muscle proteolysis requires transamination to 4-methyl-2-oxopentanoate. The inhibition of muscle protein degradation by leucine is most sensitive to, but not specific for, its 2-oxo acid, 4-methyl-2-oxopentanoate.  相似文献   

8.
Protein splicing involves the self-catalyzed excision of a protein-splicing element, the intein, from flanking polypeptides, the exteins, which are concomitantly joined by a peptide bond. Taking advantage of recently developed in vitro systems in which protein splicing occurs in trans to assay for protein-splicing inhibitors, we discovered that low concentrations of Zn(2+) inhibited splicing mediated both by the RecA intein from Mycobacterium tuberculosis and by the naturally split DnaE intein from Synechocystis sp. PCC6803. Inhibition by Zn(2+) was also observed with a cis-splicing system involving the RecA intein. In all experimental systems used, inhibition by Zn(2+) could be completely reversed by the addition of EDTA. Zinc ion also inhibited hydroxylamine-dependent N-terminal cleavage of the RecA intein. All other divalent transition metal ions tested were less effective as inhibitors than Zn(2+). The reversible inhibition by Zn(2+) should be useful in studies of the mechanism of protein splicing and allow structural studies of unmodified protein-splicing precursors.  相似文献   

9.
To evaluate the importance of RecA in DNA double-strand break (DSB) repair, we examined the effect of low and high RecA concentrations such as 2500 and 100 000 molecules per cell expressed from the inducible Pspac promoter in Deinococcus radiodurans in absence or in presence of IPTG respectively. We showed that at low concentration, RecA has a negligible effect on cell survival after gamma-irradiation when bacteria were immediately plated on TGY agar whereas it significantly decreased the survival to gamma-irradiation of DeltaddrA cells while overexpression of RecA can partially compensate the loss of DdrA protein. In contrast, when cells expressing limited concentration of RecA were allowed to recover in TGY2X liquid medium, they showed a delay in mending DSB, failed to reinitiate DNA replication and were committed to die during incubation. A deletion of irrE resulted in sensitivity to gamma-irradiation and mitomycin C treatment. Interestingly, constitutive high expression of RecA compensates partially the DeltairrE sensitization to mitomycin C. The cells with low RecA content also failed to cleave LexA after DNA damage. However, neither a deletion of the lexA gene nor the expression of a non-cleavable LexA(Ind-) mutant protein had an effect on survival or kinetics of DNA DSB repair compared with their lexA+ counterparts in recA+ as well as in bacteria expressing limiting concentration of RecA, suggesting an absence of relationship between the absence of LexA cleavage and the loss of viability or the delay in the kinetics of DSB repair. Thus, LexA protein seems to play no major role in the recovery processes after gamma-irradiation in D. radiodurans.  相似文献   

10.
The oxidative stress response in Bacillus subtilis   总被引:9,自引:0,他引:9  
Abstract Bacillus subtilis undergoes a typical bacterial stress response when exposed to low concentrations (0.1 mM) of hydrogen peroxide. Protection is thereby induced against otherwise lethal, challenge concentrations (10 mM) of this oxidant and a number of proteins are induced including the scavenging enzymes, catalase and alkyl hydroperoxide reductase, and a putative DNA binding and protecting protein. Induced protection against higher concentrations (10–30 mM) of hydrogen peroxide is eliminated in a catalase-deficient mutant. Both RecA and Spo0A influence the basal but not the induced resistance to hydrogen peroxide. A regulatory mutation has been characterized that affects the inducible phenotype and is constitutively resistant to high concentrations of hydrogen peroxide. This mutant constitutively overexpresses the proteins induced by hydrogen peroxide in the wild-type. The resistance of spores to hydrogen peroxide is partly attributable to binding of small acid soluble proteins by the spore DNA and partly to a second step which coincides with the depletion of the NADH pool, which may inhibit the generation of hydroxyl radicals from hydrogen peroxide.  相似文献   

11.
Malonaldehyde was prepared from 1,3-propanediol by alcohol dehydrogenase. The Km for 1,3-propanediol was about 1.7 mM. The reaction proceeded best at low ionic strength and at pH 9. The reaction was unaffected by pyrophosphate, phosphate, bicarbonate, or N-ethylmorpholine buffers, or by Mg+2, Ca+2, EDTA, or citrate. However, the reaction was inhibited 50% by 1.5 mM borate, 1 mM cyanide, and 5 mM azide. Thiols, such as dithioerythritol, inhibited the reaction 50% at 50–100 μM, while others, such as mercaptoacetate, inhibited 50% at concentrations over 1 mM. Malonaldehyde was removed from the reaction mixture by evaporation at pH 3 and condensation at ?78°C. No other products associated with lipid peroxidation were produced. The method was useful for preparation of radiolabeled malonaldehyde.  相似文献   

12.
We have prepared a mutant RecA protein in which proline 67 and glutamic acid 68 in the NTP binding site were replaced by a glycine and alanine residue, respectively. The [P67G/E68A]RecA protein catalyzes the single-stranded DNA-dependent hydrolysis of ATP and is able to promote the standard ATP-dependent three-strand exchange reaction between a circular bacteriophage phiX174 (phiX) single-stranded DNA molecule and a homologous linear phiX double-stranded (ds) DNA molecule (5.4 kilobase pairs). The strand exchange activity differs from that of the wild type RecA protein, however, in that it is (i) completely inhibited by an ATP regeneration system, and (ii) strongly stimulated by the addition of high concentrations of ADP to the reaction solution. These results indicate that the strand exchange activity of the [P67G/E68A]RecA protein is dependent on the presence of both ATP and ADP. The ADP dependence of the reaction is reduced or eliminated when (i) a shorter linear phiX dsDNA fragment (1.1 kilobase pairs) is substituted for the full-length linear phiX dsDNA substrate, or (ii) the Mg(2+) concentration is reduced to a level just sufficient to complex the ATP present in the reaction solution. These results indicate that it is the branch migration phase (and not the initial pairing step) of the [P67G/E68A]RecA protein-promoted strand exchange reaction that is dependent on ADP. It is likely that the [P67G/E68A]RecA mutation has revealed a requirement for ADP that also exists (but is not as readily apparent) in the strand exchange reaction of the wild type RecA protein.  相似文献   

13.
Polyamine-activated protein phosphatase activity in HeLa cell nuclei   总被引:2,自引:0,他引:2  
Protein phosphatase activity towards endogenous nuclear substrates in sonicates of isolated nuclei was activated 2-4-fold by spermine. Exogenous casein was dephosphorylated by these preparations only in the presence of spermine. Activation by spermine was half maximal at about 0.1 mM. Spermidine also activated, with half maximal stimulation at 1mM; putrescine activated poorly. Mg++ and Ca++ appeared to activate the same phosphatase activity but were only 50% as effective as spermine. Spermine activation was inhibited by 200 mM NaCl, 50 mM NaF, or 40 mM beta-glycerol phosphate. Nuclear phosphatase activity, with or without spermine, was inhibited 50% by inhibitor 2 of protein phosphatase 1. These observations suggest that protein phosphatase 1 is a major nuclear protein phosphatase and that its activity against endogenous nuclear substrates is activated by physiological concentrations of spermine.  相似文献   

14.
Monovalent cations such as Na+ and K+ inhibit the activity of T4 DNA ligase. However, the extent of inhibition varies with the terminal sequence of the duplex DNA used as substrate; in many cases, ligation of DNA is completely inhibited at 200 mM. The activity of the ligase is stimulated by raising the concentration of polyethylene glycol 6000 from 0 to 15% (w/v) when NaC1 and KC1 were both absent. Ligation was reduced as the concentration of NaC1 or KC1 was raised in a mixture containing 5 or 15% PEG 6000. With 10% PEG 6000, both cohesive- and blunt-end ligation of this ligase increased at high concentrations of salt (150-200 mM NaC1, or 200-250 mM KC1). Further, with 10% PEG 6000, inter- and intramolecular ligation occurred at low salt concentrations (0-100 mM NaC1, or 0-150 mM KC1); only linear oligomers were formed by intermolecular ligation at the high concentrations.  相似文献   

15.
The effects of ethanol on inducible prostaglandin production in RAW macrophages were investigated. Indomethacin (1 microM) or cycloheximide (1 microM) abolished prostaglandin E2 (PGE2) production induced by lipopolysaccharide (LPS, 1 microg/ml). Ethanol at concentrations from 100 mM to 600 mM concentration-dependently inhibited inducible PGE2 production, while ethanol only at higher concentrations (400 mM or more) showed cytotoxity to the cells. Cyclooxygenase-2 (COX-2) activity, estimated by transformation of exogenous arachidonic acid into PGE2, was not affected by ethanol (100-400 mM). LPS-induced expression of COX-2 mRNA was inhibited by ethanol (50-400 mM). On the other hand, protein expression of COX-2 by LPS was significantly increased by ethanol (100-400 mM). Ethanol alone at concentrations up to 600 mM did not induce expression of COX-2 protein. In a medium containing arachidonic acid (1 microM), ethanol at a low concentration (100 mM) did not significantly affect LPS-induced PGE2 production. These results suggest that ethanol shows diverse effects on the pathway of inducible PGE2 production in macrophages. Finally, ethanol may suppress utilization of arachidonic acid, resulting in reduction of inducible PGE2 production. Further study is needed to elucidate the mechanism of dissociation of ethanol effects on protein and mRNA expression.  相似文献   

16.
A partial purification of the Epstein-Barr-virus nuclear antigen 2A (EBNA 2A) protein from the Epstein-Barr-virus-infected lymphoblastoid cell line, Cherry, has been designed. The main purification step was immunoaffinity chromatography, based on the mAb, 115E, directed towards the carboxy terminus of EBNA 2A. This was followed by chromatography over a Blue Sepharose column. According to silver-stained SDS/PAGE, EBNA 2A was estimated to be 20% pure. The purified fractions contained an ATPase activity that was inhibited by the mAb 115E. Immunopurification of six EBNA-2A-positive cell lines and their negative counterpart showed that only fractions from EBNA-2A-positive lines contained ATPase activity. In gel-filtration experiments EBNA 2A eluted as a 75-kDa protein in conjunction with an ATPase activity. The EBNA 2A protein was covalently labeled by the ATP analog [14C]5'-[p-(fluorosulfonyl)benzoyl]adenosine. The ATPase activity was found to be optimal in the presence of 0.25 mM MgCl2 or CaCl2, whereas, in the presence of MnCl2 and ZnCl2, the activity was only about 50% of the control. High concentrations of Na2VO3 and heparin do not interfere with the activity, while 2.5 mM NaF or 0.5 M NaCl give a 50% reduction of the activity. The Km for ATP and for GTP was 13 microM and 11 microM, respectively, and the Vmax for ATP was about six-times higher than with GTP as substrate. Other low-molecular-mass non-protein phosphate esters, such as phosphoserine or phosphothreonine inhibited the ATPase activity with a Ki of 18 and 32 microM, respectively. Phosphotyrosine had a Ki of 480 microM. Serine, threonine and tyrosine had no inhibitory effect on the ATPase activity.  相似文献   

17.
The amidolytic activity of chymotrypsin for Suc-Ala2-Pro-Phe-MCA was somewhat enhanced by dimyristoyl PC at low ionic strength, but not at high ionic strength. The activity was strongly inhibited by pure egg yolk PA. The inhibition by 200 ng PA was neutralized by addition of 1 microgram dimyristoyl PC or pure egg yolk PC, which formed vesicles with the PA. The Km and kcat (s-1) values of chymotrypsin for hydrolysis of Suc-Ala2-Pro-Phe-MCA changed from 15 microM to 42 microM, 0.1 mM and 0.5 mM, and from 1.5 to 2.7, 3.7, and 1.0 in the presence of 1 microgram dimyristoyl PC, 0.5 micrograms pure egg yolk PE and 0.2 microgram egg yolk PA, respectively. Gel-filtration chromatography showed that dimyristoyl PC formed a complex with chymotrypsin, but did not interact with the substrate, indicating that the basic globular protein, chymotrypsin, interacted with net-neutral PL.  相似文献   

18.
Freeze-etch electron microscopy of pure RecA protein aggregates, as well as of RecA protein complexes on single-stranded and double-stranded DNA formed with various nucleotides, has permitted a clearer discrimination between the two different helical polymers that this protein forms. Both are continuous, single-start, right-handed helices; however, the form observed when ATP or non-hydrolyzable ATP analogs are present has a pitch of 9.5 nm and a diameter of 10 nm, while the other form, observed in the absence of ATP or its analogs, or in the presence of ADP, has a pitch of 6 nm and a diameter of 12 nm. The former "long pitch" helix is found only when RecA protein is bound to DNA. The latter "short pitch" helix is also observed in pure RecA protein polymers (also termed rods) and in the needle-like paracrystals of RecA protein that form in the presence of magnesium or spermidine ions, representing bundles of rods closely packed in register. Addition of ATP or non-hydrolyzable ATP analogs in the absence of DNA dissociates the pure RecA protein crystals, as well as individual helical rods, into short curvilinear chains of attached monomers. These chains typically form closed, circular rings of 7(+/- 1) protein monomers, similar in construction to a single turn of the RecA protein helix, but significantly broader in diameter. The role of ATP in interconverting the various polymeric forms of RecA protein is discussed within the context that ATP functions as a reversible allosteric effector of RecA protein, much as it mediates reversible conformational changes in other vectoral motor proteins such as myosin, dynein, kinesin and the 70,000 Mr "heat shock" ATPases. We discuss how cyclic conversions back and forth between the short- and long-pitch conformations of RecA protein could mediate in reversible single-stranded and double-stranded DNA interactions during the search for homology.  相似文献   

19.
D-GALACTOSE TRANSPORT BY SYNAPTOSOMES ISOLATED FROM RAT BRAIN   总被引:5,自引:3,他引:2  
Abstract— Synaptosomes prepared by differential and Ficoll density gradient centrifugation took up d -galactose by two saturable transport systems: one. a high affinity system with a K m of 0-25 mn and Vmax of 075 nmol/mg protein 3 min, the other, a low affinity system with a Km of 47 mM and a Vmax of 135 nmol/mg protein/3 min. The high affinity system was inhibited by 1-5 mM phlorizin but was unaffected by the absence of sodium ion or the presence of 1 mM ouabain. The low affinity system was unaffected by phlorizin or ouabain. Both systems were inhibited by high concentrations of glucose. 2-deoxyga-lactose. and inositol, and by 2.4-dinitrophcnol. Galactose was rapidly converted in synaptosomes to phos-phorylatcd intermediates and was more slowly oxidized to 14CO2  相似文献   

20.
Rad51 protein is a eukaryotic homologue of RecA protein that is essential for homologous recombination. We developed a simple procedure for purifying yeast Rad51 protein, characterized its interaction with DNA, and compared it with those of RecA from Escherichia coli and Rad51 from higher eukaryotes. Fractionation of crude extract with 0.2% polyethylenimine eliminated contaminant proteins and nucleic acids, which can perturb the subsequent purification steps. Binding of Rad51 to single-stranded DNA was detected in solution by measuring the fluorescence anisotropy of a fluorescein probe attached to the 5' end of the oligonucleotides. The interaction was stabilized by ATP, as is that of RecA, but was neither stabilized by a non-hydrolysable analog of ATP, nor destabilized by ADP, unlike the interaction of RecA. This character was very similar to that of Xenopus XRad51.1, although the binding of yeast Rad51 to DNA was more sensitive to Mg(2+) ion in both the presence and absence of ATP, and was optimal at 5--10 mM Mg(2+). The dissociation of Rad51 protein from DNA is not, therefore, favored by the hydrolysis of ATP to ADP, in contrast to that of RecA. On the other hand, the high DNA-binding state of the Rad51-DNA complex promoted by ATP appeared to be short-lived. These features may be linked to the lower activity of Rad51 and the fact that Rad51 activity does not require the hydrolysis of ATP.  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