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.
J W Lee  M M Cox 《Biochemistry》1990,29(33):7677-7683
There are at least two major conformations of recA nucleoprotein filaments formed on poly-(deoxythymidylic acid) [poly(dT)], one stabilized by ATP [or adenosine 5'-O-(3-thiotriphosphate) (ATP gamma S)] and one stabilized by ADP. Assembly of filaments in the ATP conformation is much faster than assembly in the ADP conformation. A third conformation may be present in the absence of nucleotides. The ATP and ADP conformations are mutually exclusive. When a mixture of ATP and ADP is present, recA protein binding is a function of the ADP/ATP ratio. Complete dissociation is observed when the ratio becomes 1.0-1.5. When a mixture of ATP and ADP is present at the beginning of a reaction, a transient phase lasting several minutes is observed in which the system approaches the state characteristic of the new ADP/ATP ratio. This phase is manifested by a lag in ATP hydrolysis when ATP is added to preformed ADP filaments, and by a burst in ATP hydrolysis in all other cases. More than 15 ATPs are hydrolyzed per bound recA monomer during the burst phase. The transient phase reflects an end-dependent disassembly process propagated longitudinally through the filament, rather than a slow conformation change in individual recA monomers or a slow exchange of one nucleotide for the other. The hysteresis exhibited by the system provides a number of insights relevant to the mechanism of recA-mediated DNA strand exchange.  相似文献   

2.
The interaction of recA protein with single-stranded (ss) phi X174 DNA has been examined by means of a nuclease protection assay. The stoichiometry of protection was found to be 1 recA monomer/approximately 4 nucleotides of ssDNA both in the absence of a nucleotide cofactor and in the presence of ATP. In contrast, in the presence of adenosine 5'-O-(thiotriphosphate) (ATP gamma S) the stoichiometry was 1 recA monomer/approximately 8 nucleotides. No protection was seen with ADP. In the absence of a nucleotide cofactor, the binding of recA protein to ssDNA was quite stable as judged by equilibration with a challenge DNA (t1/2 approximately 30 min). Addition of ATP stimulated this transfer (t1/2 approximately 3 min) as did ADP (t1/2 approximately 0.2 min). ATP gamma S greatly reduced the rate of equilibration (t1/2 greater than 12 h). Direct visualization of recA X ssDNA complexes at subsaturating recA protein concentrations using electron microscopy revealed individual ssDNA molecules partially covered with recA protein which were converted to highly condensed networks upon addition of ATP gamma S. These results have led to a general model for the interaction of recA protein with ssDNA.  相似文献   

3.
J W Lee  M M Cox 《Biochemistry》1990,29(33):7666-7676
ADP and adenosine 5'-O-(3-thiotriphosphate) (ATP gamma S) inhibit recA protein promoted ATP hydrolysis by fundamentally different mechanisms. In both cases, at least two modes of inhibition are observed. For ADP, the first mode is competitive inhibition. The second mode is manifested by dissociation of recA protein from DNA. These are readily distinguished in a comparison of ATP hydrolyses that are activated by (a) DNA and (b) high (approximately 2 M) salt concentrations. Competitive inhibition with a significant degree of cooperativity is observed under both sets of conditions, although the DNA-dependent activity is more sensitive to ADP than the high-salt reaction. The reaction in the presence of poly(deoxythymidylic acid) or duplex DNA ceases when about 60% of the available ATP is hydrolyzed, reflecting an ADP-mediated dissociation of recA protein from the DNA that is governed by the ADP/ATP ratio. In contrast, ATP hydrolysis proceeds nearly to completion at high salt concentrations. At high concentrations of ATP and ATP gamma S, ATP gamma S also acts as a competitive inhibitor. At low concentrations of ATP gamma S and ATP, however, ATP gamma S activates ATP hydrolysis. These patterns are observed for recA-mediated ATP hydrolysis with either high salt concentrations or a poly(deoxythymidylic acid) [poly(dT)] cofactor, although the activation is observed at much lower ATP and ATP gamma S concentrations when poly(dT) is used. ATP gamma S can also relieve the inhibitory effect of ADP under some conditions. ATP gamma S and ADP are antagonistic inhibitors, reinforcing the idea that they stabilize different conformations of the protein and suggesting that these conformations are mutually exclusive. The ATP gamma S (ATP) conformation is active in ATP hydrolysis. The ADP conformation is inactive.  相似文献   

4.
The transfer of recA protein from a fluorescently modified single-stranded DNA, containing 1,N6-ethenoadenosine and 3,N4-ethenocytosine, to polydeoxythymidylic acid (poly(dT)) was shown to occur by a complex mechanism in both the absence and presence of ADP (Menetski, J. P., and Kowalczykowski, S. C. (1987) J. Biol. Chem. 262, 2085-2092). A part of the mechanism involves the formation of a kinetic ternary intermediate. Since the binding and hydrolysis of ATP by recA protein is involved in many of the recA protein in vitro activities, we have analyzed the effect of ATP on the transfer reaction. In the presence of ATP, the transfer reaction is dependent on the concentration of the competitor single-stranded DNA, poly(dT). This result suggests that transfer does not occur by a simple dissociation mechanism. The reaction occurs via two kinetically distinct species of protein X DNA complexes with properties that are similar to those characterized for the transfer reaction in the absence of ATP. There is a complicated effect of nucleotide concentration on the rate of transfer. At low concentrations of ATP (less than 50 microM), increasing nucleotide concentration increases the rate of transfer; this is similar to the effect of ADP. However, at high concentrations of ATP (greater than 50 microM), increasing ATP concentration decreases the rate of transfer. Finally, the processivity of ATP hydrolysis during transfer was found to increase with increases in ATP concentration. Less than one ATP molecule was hydrolyzed per transfer event at low ATP concentrations (less than 20 microM) while greater than 50 molecules were hydrolyzed at high ATP concentration (greater than 250 microM). These data suggest that the rate of transfer is not directly coupled to the rate of hydrolysis.  相似文献   

5.
M S Silver  A R Fersht 《Biochemistry》1983,22(12):2860-2866
The availability of epsilon DNA, a fluorescent ssDNA derivative, has made it possible to examine quantitatively the interactions between recA protein and single-stranded polynucleotides. Fluorescence titrations of epsilon DNA with recA protein and vice versa establish that each recA protein monomer covers 5.5 epsilon DNA nucleotides and that the dissociation constant of the recA-epsilon DNA complex is 10 nM. Fluorescence titrations of recA protein-epsilon DNA mixtures with poly(dT) establish that each recA protein monomer covers 5.1 poly(dT) nucleotides and that the dissociation constant of the recA-poly(dT) complex is 0.03 nM. Observations on how the addition of ssDNA affects the fluorescence of recA protein-epsilon DNA mixtures establish that the dissociation constant of the recA-ssDNA complex exceeds 20 microM. Stopped-flow kinetics in which excess recA protein binds to epsilon DNA indicate that k2 = 6 X 10(6) M-1 s-1 for the process. A more approximate kinetic technique indicates that recA protein binds to epsilon DNA at least one-tenth as fast as to poly(dT); the rate constant for dissociation of recA-epsilon DNA exceeds that for recA-poly(dT) by at least 30-fold. epsilon DNA is proven to be a versatile reagent for studying single-stranded polynucleotide-protein interactions. Not only can its own complexes with protein be investigated but also, under suitable circumstances, it can be used as a fluorescent probe to explore complexes incorporating nonfluorescent polynucleotides.  相似文献   

6.
When recA protein binds cooperatively to single-stranded DNA to form filamentous nucleoprotein complexes, it becomes competent to hydrolyze ATP. No correlation exists between the ends of such complexes and the rate of ATP hydrolysis. ATP hydrolysis is not, therefore, restricted to the terminal subunits on cooperatively bound recA oligomers, but occurs throughout the complex. Similarly, during recA protein-promoted branch migration (during DNA strand exchange), ATP hydrolysis is not restricted to recA protein monomers at the branch point. DNA cofactors of lengths varying from 16 bases to over 12,000 bases support ATP hydrolysis. The maximum value of kcat at infinite DNA concentration is about 29/min independent of the length of the DNA cofactor. The apparent dissociation constant, however, is a strong function of DNA length, providing evidence for a minimum site size of 30-50 bases for efficient binding of recA protein.  相似文献   

7.
Structure of the RecA x ADP(ATP) and recA x ADP x cation(+2) complexes was studied by methods of ESR, NMR and near-ultraviolet spectroscopy. The strong hypochromism in the adenine absorption band occurs. The complexes of nucleotide with cation and with protein were independently involved in the triple recA x ADP x cation(+2) complex. The triple complex can be treated as a three-link chain with the ADP localized in the middle.  相似文献   

8.
We have examined the exchange of recA protein between stable complexes formed with single-stranded DNA (ssDNA) and (a) other complexes and (b) a pool of free recA protein. We have also examined the relationship of ATP hydrolysis to these exchange reactions. Exchange was observed between two different recA X ssDNA complexes in the presence of ATP. Complete equilibration between two sets of complexes occurred with a t1/2 of 3-7 min under a set of conditions previously found to be optimal for recA protein-promoted DNA strand exchange. Approximately 200 ATPs were hydrolyzed for every detected migration of a recA monomer from one complex to another. This exchange occurred primarily between adjacent complexes, however. Little or no exchange was observed between recA X ssDNA complexes and the free recA protein pool, even after several hundred molecules of ATP had been hydrolyzed for every recA monomer present. ATP hydrolysis is not coupled to complete dissociation or association of recA protein from or with recA X ssDNA complexes under these conditions.  相似文献   

9.
We have investigated quantitative molecular aspects of the interaction of recA protein with single-stranded DNA, by using a fluorescent modified-DNA referred to as etheno-M13 DNA. In addition, the effects of the nucleotide cofactors ATP and ADP, and the analogues ATP-gamma-S, AMP-P-C-P, and AMP-P-N-P on this interaction have been studied. It is shown that ATP, AMP-P-N-P and, in particular, ATP-gamma-S significantly increase the affinity of recA protein for single-stranded DNA, whereas ADP and, to a lesser degree, AMP-P-C-P decrease the affinity. Binding to etheno-M13 single-stranded DNA is co-operative, with the value of the co-operativity parameter, omega, being approximately 50 under all conditions measured. The effect that ADP has on recA protein-DNA affinity is to lower the intrinsic binding constant, but it has no effect on the co-operativity of binding. In addition, the stability of the recA protein-DNA complex is very salt dependent (d log K/d log [NaC1] approximately -10) and it is the intrinsic binding affinity rather than the co-operativity of binding that is affected; thus, under all conditions observed, recA protein binds single-stranded DNA co-operatively with a value of omega = 50 +/- 10. The binding affinity is also influenced by the type of anion present, being approximately 10,000-fold higher when acetate ion is present instead of chloride ion. These data have been interpreted to suggest that recA protein forms up to five ionic interactions when it binds to single-stranded DNA and that five to six anions are displaced upon binding. The modulation of recA protein-DNA complex stability by nucleotide cofactors suggests that these cofactors play a role in the cycling of recA protein on and off single-stranded DNA, with ATP being required for DNA binding under physiological conditions and ADP serving as a "release" factor. These results are discussed in terms of a model for the role of ATP hydrolysis in a recA protein-single stranded DNA binding cycle.  相似文献   

10.
The complete exchange of strands between circular single-stranded and full length linear duplex DNAs promoted by the recA protein of Escherichia coli is dependent upon the hydrolysis of ATP and is strongly stimulated by the single-stranded DNA binding protein (SSB). In the presence of SSB, stable complexes of recA protein and single-stranded DNA are formed as an early step in the reaction. These complexes dissociate when the ADP/ATP ratio approaches a value of 0.6-1.5, depending upon reaction conditions. Thus, ATP hydrolysis never proceeds to completion but stops when 40-60% of the input ATP has undergone hydrolysis. recA protein can participate in a second round of strand exchange upon regeneration of the ATP. While 100-200 mol of ATP are hydrolyzed/mol of heteroduplex base pair formed under standard reaction conditions in the presence of SSB, this value is reduced to 16 at levels of ADP lower than that required to dissociate the complexes. ATP hydrolysis appears to be completely irreversible since efforts to detect exchange reactions using 18O probes have been unsuccessful.  相似文献   

11.
In an attempt to understand the role of ATP as a cofactor at the interaction of the RecA protein with DNA, we have studied the orientation geometries of the cofactor analogs adenosine 5′-O-(3-thiotriphosphate) (ATPγS) and guanosine 5′-O-(3-thiotriphosphate) (GTPγS) in RecA-DNA complexes using flow linear dichroism spectroscopy. Both cofactors promote the formation of RecA-DNA complexes of similar structure as judged from similar orientations of DNA bases. The DNA orientation was probed through the dichroism of the long-wavelength absorption of a DNA analog, poly(dεA). In this way differences between the dichroic spectra of the ATPγS–RecA–DNA and GTPγS-RecA-DNA complexes, observed in the shorter-wavelength region, are related to orientation at variations of the cofactor chromophores. The results show that the guanine plane of GTPγS is oriented parallel with the principal axis of the complex in contrast to the more perpendicular orientation of the DNA bases. This observation directly excludes the possibility that the cofactor could be intercalated between the DNA bases. This observation directly excludes the possibility that the cofactor could be intercalated between the DNA bases. The orientation of the adenine base of ATPγS, which may be similar to that of guanine of GTPγS albeit not exactly the same, is also inconsistent with intercalation. The possibility that the cofactor bound to the protein could be intercalated in DNA had been speculated from the observation that some DNA intercalators can induce RecA binding to DNA in the absence of cofator. There are probably no direct interactions between the cofator and the DNA bases and the role of the cofactor is probably related to interaction with RecA and a modification of protein conformation.  相似文献   

12.
Binding of the recA gene product from Escherichia coli to single-stranded polynucleotides has been investigated using poly(dA) that have been modified by chloroacetaldehyde to yield fluorescent 1,N6-ethenoadenine (epsilon A) bases. A strong enhancement of the fluorescent quantum yield of poly(d epsilon A) is induced upon RecA protein binding. A 4-fold increase is observed in the absence of ATP or ATP gamma S and a 7-fold increase in the presence of either nucleoside triphosphate. RecA protein can bind to poly(d epsilon A) in the absence of both Mg2+ ions and ATP (or ATP gamma S) but Mg2+ ions are required to observe RecA protein binding in the presence of ATP (or ATP gamma S) at pH 7.5. ATP binding to the RecA-poly(d epsilon A) complex induces a dissociation of RecA from the polynucleotide followed by re-binding of [RecA-ATP-Mg2+] ternary complex. Whereas ATP-induced dissociation of RecA-poly(d epsilon A) complexes is a fast process, the subsequent binding reaction of [RecA-ATP-Mg2+] is slow. A model is proposed whereby [RecA-ATP-Mg2+] binding to poly(d epsilon A) involves slow nucleation and elongation processes along the polynucleotide backbone. The nucleation reaction is shown to involve at least a trimer or a tetramer. Polymerization of the [RecA-ATP-Mg2+] ternary complex stops when the polynucleotide is entirely covered with 6 +/- 1 nucleotides per RecA monomer. ATP hydrolysis then induces a release of RecA-ADP complexes from the polynucleotide template.  相似文献   

13.
We have analyzed the transfer kinetics of recA protein from one polynucleotide to another by monitoring the change in fluorescence of a modified single-stranded M13 DNA, referred to as etheno M13 DNA, that accompanies recA protein dissociation. The observed rate of transfer is dependent on the concentration of competitor polynucleotide, polythymidylic acid (poly(dT]; increasing the poly(dT) concentration increases the rate of transfer. These data are inconsistent with the rate-limiting step in the transfer mechanism occurring by a simple dissociation process. Under certain conditions, the apparent rate constant displays plateau behavior at high poly(dT) concentrations. This result is indicative of transfer occurring through a ternary intermediate including etheno M13 DNA and poly(dT). The transfer reaction was found to occur through two kinetically distinct species of transferring recA protein X DNA complexes. The relative amounts of these two species was affected by both the MgCl2 and protein concentration, suggesting that the two kinetic components reflect different aggregation states of the recA protein X DNA complex. Because etheno M13 DNA and poly(dT) contain no complementary sequences, we have concluded that recA protein has the intrinsic ability to form a kinetic ternary intermediate with two separate single-stranded DNA molecules in the absence of homology.  相似文献   

14.
We investigated the use of fluorescence lifetime imaging microscopy (FLIM) of a fluorescently labeled ATP analog (3'-O-{N-[3-(7-diethylaminocoumarin-3-carboxamido)propyl]carbamoyl}ATP) to probe in permeabilized muscle fibers the changes in the environment of the nucleotide binding pocket caused by interaction with actin. Spatial averaging of FLIM data of muscle sarcomeres reduces photon noise, permitting detailed analysis of the fluorescence decay profiles. FLIM reveals that the lifetime of the nucleotide, in its ADP form because of the low concentration of nucleotide present, changes depending on whether the nucleotide is free in solution or bound to myosin, and on whether the myosin is bound to actin in an actomyosin complex. Characterization of the fluorescence decays by a multiexponential function allowed us to resolve the lifetimes and amplitudes of each of these populations, namely, the fluorophore bound to myosin, bound to actin, in an actomyosin complex, and free in the filament lattice. This novel application of FLIM to muscle fibers shows that with spatial averaging, detailed information about the nature of nucleotide complexes can be derived.  相似文献   

15.
16.
A soluble form of CD39 was expressed and purified from High-Five insect cells. The soluble CD39 is a monomer with a molecular weight of 54,000. The k(cat) and K(m) of the purified soluble CD39 were 4.6 s(-1) and 12 microM for ATP and 1.3 s(-1) and 7 microM for ADP as substrates, respectively. One nucleotide binding site was detected on the monomer only in the presence of Ca(2+). In contrast to the membrane bound CD39, soluble CD39 released ADP as an intermediate during ATP hydrolysis, as did the soluble potato apyrase.  相似文献   

17.
The effects of adenine nucleotides on pea seed glutamine synthetase (EC 6.3.1.2) activity were examined as a part of our investigation of the regulation of this octameric plant enzyme. Saturation curves for glutamine synthetase activity versus ATP with ADP as the changing fixed inhibitor were not hyperbolic; greater apparent Vmax values were observed in the presence of added ADP than the Vmax observed in the absence of ADP. Hill plots of data with ADP present curved upward and crossed the plot with no added ADP. The stoichiometry of adenine nucleotide binding to glutamine synthetase was examined. Two molecules of [gamma-32P]ATP were bound per subunit in the presence of methionine sulfoximine. These ATP molecules were bound at an allosteric site and at the active site. One molecule of either [gamma-32P]ATP or [14C]ADP bound per subunit in the absence of methionine sulfoximine; this nucleotide was bound at an allosteric site. ADP and ATP compete for binding at the allosteric site, although ADP was preferred. ADP binding to the allosteric site proceeded in two kinetic phases. A Vmax value of 1.55 units/mg was measured for glutamine synthetase with one ADP tightly bound per enzyme subunit; a Vmax value of 0.8 unit/mg was measured for enzyme with no adenine nucleotide bound at the allosteric site. The enzyme activation caused by the binding of ADP to the allosteric sites was preceded by a lag phase, the length of which was dependent on the ADP concentration. Enzyme incubated in 10 mM ADP bound approximately 4 mol of ADP/mol of native enzyme before activation was observed; the activation was complete when 7-8 mol of ADP were bound per mol of the octameric, native enzyme. The Km for ATP (2 mM) was not changed by ADP binding to the allosteric sites. ADP was a simple competitive inhibitor (Ki = 0.05 mM) of ATP for glutamine synthetase with eight molecules of ADP tightly bound to the allosteric sites of the octamer. Binding of ATP to the allosteric sites led to marked inhibition.  相似文献   

18.
Glycerol-extracted rabbit psoas fibres were incubated at temperatures between -35 degrees C and +10 degrees C in a low-ionic-strength relaxing solution containing 50% ethyleneglycol, 100 microM [3H]MgATP, 1 mM [14C]mannitol and less than 0.01 microM Ca2+. The fibres were then rinsed in a solution containing 1 mM ATP and the bound nucleotide eluted in trichloroacetic acid; all these operations were carried out at the cold temperature. Residual bound nucleotide was eluted with trichloroacetic acid at room temperature. The fibres were found to bind approximately 180 microM nucleotide, which is consistent with binding to the enzymatic site of myosin. The eluate, obtained in the cold, was analysed on poly(ethyleneimine)-cellulose for its ATP and ADP content. At temperatures down to -22 degrees C most of the bound nucleotide was ADP and there was little variation of this fraction with temperature. As the temperature was lowered below -22 degrees C the ATP fraction rose sharply; by -35 degrees C it predominated. These results are similar in type to those found by Biosca et al. [(1984) Biochemistry 23, 1947-1953] on isolated subfragment 1, but are displaced to a much lower temperature range. Thus in a muscle fibre only a low thermal energy is needed for myosin to hold its nucleotide in a constant balance between ATP and ADP.  相似文献   

19.
Left-handed Z-DNA binding by the recA protein of Escherichia coli   总被引:10,自引:0,他引:10  
recA binding to left-handed Z-DNA was measured using nitrocellulose filter binding assays with four DNA polymers with defined nucleotide sequences and four recombinant plasmids. Two to 7-fold preferential binding of recA to Z-DNA polymers was observed. Left-handed Z-DNA polymer binding by recA required ATP or its nonhydrolyzable analog, ATP(gamma S), while ADP inhibited binding. Complex formation with both B- and Z-forms was influenced by polymer length; recA bound longer DNAs better. recA binding to recombinant plasmids containing supercoil-stabilized Z-DNA was essentially similar to that found for the control vector; thus, no preferential binding of recA to the Z-form was observed. Comparative experiments with the rec1 protein of Ustilago maydis and the Escherichia coli recA protein were performed. In our hands, recA and rec1 have a similar capacity for binding left-handed Z-DNA polymers and for binding recombinant plasmids containing B- and/or Z-regions. recA contains a left-handed Z-DNA-stimulated ATPase activity. This activity differs from the right-handed B-DNA-stimulated activity since it is less sensitive to increasing pH. The kinetics of ATP hydrolysis in B-DNA/Z-DNA mixing experiments showed that the turnover of the Z-DNA recA complex was slower than for B-DNA suggesting that left-handed Z-DNA is more stably bound by recA. Our results are consistent with the postulate that left-handed Z-DNA is involved in genetic recombination.  相似文献   

20.
In an effort to clarify the requirement for ATP in the recA protein-promoted renaturation of complementary DNA strands, we have analyzed the mutant recA1 protein which lacks single-stranded DNA-dependent ATPase activity at pH 7.5. Like the wild type, the recA1 protein binds to single-stranded DNA with a stoichiometry of one monomer per approximately four nucleotides. However, unlike the wild type, the mutant protein is dissociated from single-stranded DNA in the presence of ATP or ADP. The ATP analogue adenosine 5'-O-3' (thiotriphosphate) appears to stabilize the binding of recA1 protein to single-stranded DNA but does not elicit the stoichiometry of 1 monomer/8 nucleotides or the formation of highly condensed protein-DNA networks that are characteristic of the wild type recA protein in the presence of this analogue. The recA1 protein does not catalyze DNA renaturation in the presence of ATP, consistent with the dissociation of recA1 protein from single-stranded DNA under these conditions. However, it does promote a pattern of Mg2+-dependent renaturation identical to that found for wild type recA protein.  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