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.
In the presence of RecA single-stranded DNA (ssDNA) is found to exhibit flow linear dichroism (LD). In the absence of the cofactor ATP gamma S, the LD is positive with a maximum at about 280 nm, whereas in the presence of the cofactor ATP gamma S there is still a positive long-wavelength band, but a negative LD contribution centered at 260 nm indicates an orientation of the DNA bases preferentially perpendicular to the fiber axis. For the complex between ssDNA and RecA without ATP gamma S, essentially all LD derives from the protein (tryptophane) subunits indicating a structure in which the tryptophanes are preferentially parallel to the fiber axis of the complex while the DNA bases remain essentially unoriented. The magnitude of the LD increases with the RecA/DNA ratio to a point corresponding to approximately three nucleotides per RecA and decreases thereafter with excess of DNA. This indicates that there are two modes of binding with different stoichiometries.  相似文献   

2.
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   

3.
We have analyzed the nature of RecA protein-RecA protein interactions using an affinity column prepared by coupling RecA protein to an agarose support. When radiolabeled soluble proteins from Escherichia coli are applied to this column, only the labeled RecA protein from the extract was selectively retained and bound tightly to the affinity column. Efficient binding of purified 35S-labeled RecA protein required Mg2+, and high salt did not interfere with the binding of RecA protein to the column. Complete removal of the bound enzyme from the affinity column required treatment with guanidine HCl (5 M) or urea (8 M). These and other properties suggest that hydrophobic interactions contribute significantly to RecA protein subunit recognition in solution. Using a series of truncated RecA proteins synthesized in vitro, we have obtained evidence that at least some of the sequences involved in protein recognition are localized within the first 90 amino-terminal residues of the protein. Based on the observation that RecA proteins from three heterologous bacteria are specifically retained on the E. coli RecA affinity column, it is likely that this binding domain is highly conserved and is required for interaction and association of RecA protein monomers. Stable ternary complexes of RecA protein and single-stranded DNA were formed in the presence of the nonhydrolyzable ATP analog adenosine 5'-O-(thiotriphosphate) and applied to the affinity columns. Most of the complexes formed with M13 DNA could be eluted in high salt, whereas a substantial fraction of those formed with the oligonucleotide (dT)25-30 remained bound in high salt and were quantitatively eluted with guanidine HCl (5 M). The different binding properties of these RecA protein-DNA complexes likely reflect differences in the availability of a hydrophobic surface on RecA protein when it is bound to long polynucleotides compared to short oligonucleotides.  相似文献   

4.
Solution structure of complexes between DNA and recombinase RecA from Escherchia coli, in the presence of the physiological cofactor ATP, is probed by flow linear dichroism (LD) spectroscopy. A problem of ADP accumulation which promotes dissociation of DNA-RecA is circumvented by using an ATP-regenerating system. The LD features indicate that the local structure of the complex is very similar to that found in the presence of the non-hydrolysable analog of ATP, adenosine-5'-O-[gamma-thio]triphosphate (ATP[gamma S]); the DNA bases are oriented with their planes preferentially perpendicular to the long axis of the filament, while the indole chromophores of the two tryptophan residues of RecA are rather parallel to this reference direction. A much smaller overall amplitude of the LD spectrum, compared to ATP[gamma S], is interpreted as a result of fast dissociation of RecA due to hydrolysis of ATP, producing transiently naked DNA regions which act like flexible joints, diminishing the macroscopic orientation of the RecA filaments. However, the ATP hydrolysis is not found to prevent simultaneous accommodation of two non-complementary DNA molecules in the RecA complex, as judged from the LD behaviour upon successive addition of two different polynucleotides or modified DNA strands. A notable difference from corresponding complexes formed with ATP[gamma S] is that, in the presence of ATP hydrolysis, the order in which the two DNA molecules have been added is insignificant as judged from virtually identical resulting structures; this observation indicates that exchange of DNA occurs between the two DNA accommodation sites within the RecA filament.  相似文献   

5.
The efficiency of single-stranded (ss) oligonucleotides binding at the secondary site of the RecA protein filament is demonstrated to depend on the strandedness of DNA bound at the primary site. When the primary site is occupied by a ss-oligonucleotide, the binding of another ss-oligonucleotide at the secondary site is characterized by higher affinity and a lower rate of dissociation than is the case when the primary site is occupied by a double-stranded oligonucleotide. In contrast to a DNA strand exchange reaction suppressed by a heterologous oligonucleotide bound at the secondary site of the RecA filament, the occupation of the secondary site by a heterologous oligonucleotide does not prevent renaturation between the oligonucleotides bound at the primary site and complementary oligonucleotides from solution demonstrating that the binding of a DNA strand in the secondary site is not a necessary intermediate step in RecA-promoted DNA renaturation.  相似文献   

6.
Isoelectric focusing (IEF) in thin capillaries is reviewed here. After an introduction on the genesis and chemistry of the carrier ampholyte buffers, different approaches to IEF are discussed and evaluated. The classical approach consists on IEF under conditions of suppressed electroosmotic (EOF) flow, usually obtained by covalently bonding hydrophilic polymers to the inner capillary wall. The other approach consists of IEF in dynamically (and partially) coated capillaries, so as to allow a reduced EOF flow to coexist with the IEF process, so that focusing and transport of the train of stacked bands occurs simultaneously. The various experimental parameters: focusing, elution and detection steps, pI measurements, as well as typical drawbacks, such as isoelectric precipitation are evaluated. The review ends with some examples of analytical separations, at the moment mostlyl limited to focusing of native hemoglobins (normal and point mutants). These separations are compared with those obtained by slab-gel IEF and in immobilized pH gradients.  相似文献   

7.
recA protein (RecA) performs diverse catalytic activities that require a complex with single-stranded DNA and an NTP. A subset of these functions shows optimal activity at a high DNA/protein ratio and requires NTP hydrolysis, whereas other catalytic activities are optimal in RecA-saturated complexes that require NTP, but do not hydrolyze it. To analyze the mechanism of catalytic discrimination, we investigated the properties of RecA bound to small oligonucleotides (oligos) of defined sizes. We show that RecA bound to (dT)16 is optimal for co-protease activity and not active as ATPase whereas the complex with (dT)24 is competent in ATP hydrolysis but impaired as a co-protease. Thermodynamic measurements of the equilibrium-binding properties of these complexes showed that (dT)24 promoted a more salt sensitive complex than the one formed with (dT)16, indicating more ionic interactions between RecA and DNA in the former. X-ray pictures show that the oligo complexes form helixes. We propose that RecA may change its conformation as a function of the number of phosphates available to the monomer in the interacting DNA lattice, thus promoting an allosteric change in catalytic activities. This model offers explanations for the observed inhibition of co-protease activity by excess ssDNA.  相似文献   

8.
The RecA protein of Escherichia coli binds specifically to acidic phospholipids such as cardiolipin and phosphatidylglycerol. This binding appears to be affected by the presence of divalent cations such as Ca2+ and Mg2+. The interaction leads to the inhibition of RecA binding to at least two different conformations of DNA, single-stranded DNA and left-handed Z-DNA, thus suggesting that the phospholipids interact at the DNA-binding site of the RecA protein. Inclusion of a nucleotide cofactor [adenosine 5'-O-(gamma-thiotriphosphate)] in the reactions did not prevent the inhibition of DNA-binding activities of RecA protein by the phospholipids. The interaction of RecA protein with cardiolipin and phosphatidylglycerol, which represent two of the three major phospholipids of the E. coli membrane, may be physiologically important, as it provides a possible mechanism for the RecA-membrane association during the SOS response. These observations raise the possibility that the Z-DNA-binding activity of RecA protein is merely a manifestation of its phospholipid-binding property.  相似文献   

9.
J M Bork  M M Cox  R B Inman 《The EMBO journal》2001,20(24):7313-7322
The Escherichia coli RecF, RecO and RecR pro teins have previously been implicated in bacterial recombinational DNA repair at DNA gaps. The RecOR-facilitated binding of RecA protein to single-stranded DNA (ssDNA) that is bound by single-stranded DNA-binding protein (SSB) is much faster if the ssDNA is linear, suggesting that a DNA end (rather than a gap) facilitates binding. In addition, the RecOR complex facilitates RecA protein-mediated D-loop formation at the 5' ends of linear ssDNAs. RecR protein remains associated with the RecA filament and its continued presence is required to prevent filament disassembly. RecF protein competes with RecO protein for RecR protein association and its addition destabilizes RecAOR filaments. An enhanced function of the RecO and RecR proteins can thus be seen in vitro at the 5' ends of linear ssDNA that is not as evident in DNA gaps. This function is countered by the RecF/RecO competition for association with the RecR protein.  相似文献   

10.
Gamper HB  Nulf CJ  Corey DR  Kmiec EB 《Biochemistry》2003,42(9):2643-2655
RecA protein catalyzes strand exchange between homologous single-stranded and double-stranded DNAs. In the presence of ATPgammaS, the post-strand exchange synaptic complex is a stable end product that can be studied. Here we ask whether such complexes can hybridize to or exchange with DNA, 2'-OMe RNA, PNA, or LNA oligonucleotides. Using a gel mobility shift assay, we show that the displaced strand of a 45 bp synaptic complex can hybridize to complementary oligonucleotides with different backbones to form a four-stranded (double D-loop) joint that survives removal of the RecA protein. This hybridization reaction, which confirms the single-stranded character of the displaced strand in a synaptic complex, might initiate recombination-dependent DNA replication if it occurs in vivo. We also show that either strand of the heteroduplex in a 30 bp synaptic complex can be replaced with a homologous DNA oligonucleotide in a strand exchange reaction that is mediated by the RecA filament. Consistent with the important role that deoxyribose plays in strand exchange, oligonucleotides with non-DNA backbones did not participate in this reaction. The hybridization and strand exchange reactions reported here demonstrate that short synaptic complexes are dynamic structures even in the presence of ATPgammaS.  相似文献   

11.
Binding of RecA to poly(dG-m5dC) and poly(dG-dC) under B- and Z-form conditions was studied using circular dichroism (CD) and linear dichroism (LD). LD revealed a quantitative binding of RecA to Mg2+-induced Z-form poly(dG-m5dC) with a stoichiometry of 3.1 base pairs/RecA monomer, which is slightly larger than the 2.7 base pairs observed for the B-form. The LD spectra indicate a preferentially perpendicular orientation of DNA bases and a rather parallel orientation of the tryptophan residues relative to the fiber axis in both complexes. The association rate of RecA to Z-form DNA was found to be slower than to B-form. CD measurements showed that the polynucleotide conformation is retained upon RecA binding, and CD and LD confirm that RecA binds to both forms of DNA. The Mg2+-induced Z-form is shown to be retransformed into B-form, both in free and in RecA-complexed polynucleotides by addition of NaCl, whereas the B----Z transition cannot be induced by addition of Mg2+ when the polynucleotide is complexed with RecA. From this it is inferred that RecA does not stabilize the Z-conformation of the polynucleotide but that it can kinetically "freeze" the polynucleotide in its B-conformation. On all essential points, the same conclusions were also reached in a corresponding study of unmethylated poly(dG-dC) with the Z-form induced by Mn2+.  相似文献   

12.
RecA protein recognises two complementary DNA strands for homologous recombination. To gain insight into the molecular mechanism, the thermodynamic parameters of the DNA binding have been characterised by isothermal calorimetry. Specifically, conformational changes of protein and DNA were searched for by measuring variations in enthalpy change (DeltaH) with temperature (heat capacity change, DeltaC(p)). In the presence of the ATP analogue ATPgammaS, the DeltaH for the binding of the first DNA strand depends upon temperature (large DeltaC(p)) and the type of buffer, in a way that is consistent with the organisation of disordered parts and the protonation of RecA upon complex formation. In contrast, the binding of the second DNA strand occurs without any pronounced DeltaC(p), indicating the absence of further reorganisation of the RecA-DNA filament. In agreement with these findings, a significant change in the CD spectrum of RecA was observed only upon the binding of the first DNA strand. In the absence of nucleotide cofactor, the DeltaH of DNA binding is almost independent of temperature, indicating a requirement for ATP in the reorganisation of RecA. When the second DNA strand is complementary to the first, the DeltaH is larger than that for non-complementary DNA strand, but less than the DeltaH of the annealing of the complementary DNA without RecA. This small DeltaH could reflect a weak binding that may facilitate the dissociation of only partly complementary DNA and thus speed the search for complementary DNA. The DeltaH of binding DNA sequences displaying strong base-base stacking is small for both the first and second binding DNA strand, suggesting that the second is also stretched upon interaction with RecA. These results support the proposal that the RecA protein restructures DNA, preparing it for the recognition of a complementary second DNA strand, and that the recognition is due mainly to direct base-base contacts between DNA strands.  相似文献   

13.
Escherichia coli dinD is an SOS gene up-regulated in response to DNA damage. We find that the purified DinD protein is a novel inhibitor of RecA-mediated DNA strand exchange activities. Most modulators of RecA protein activity act by controlling the amount of RecA protein bound to single-stranded DNA by affecting either the loading of RecA protein onto DNA or the disassembly of RecA nucleoprotein filaments bound to single-stranded DNA. The DinD protein, however, acts postsynaptically to inhibit RecA during an on-going DNA strand exchange, likely through the disassembly of RecA filaments. DinD protein does not affect RecA single-stranded DNA filaments but efficiently disassembles RecA when bound to two or more DNA strands, effectively halting RecA-mediated branch migration. By utilizing a nonspecific duplex DNA-binding protein, YebG, we show that the DinD effect is not simply due to duplex DNA sequestration. We present a model suggesting that the negative effects of DinD protein are targeted to a specific conformational state of the RecA protein and discuss the potential role of DinD protein in the regulation of recombinational DNA repair.  相似文献   

14.
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   

15.
Gamper HB  Hou YM  Kmiec EB 《Biochemistry》2000,39(49):15272-15281
Strand exchange between two duplexes is usually initiated as a three-strand event that requires the presence of a single-stranded overhang or gap in one of the two molecules. Here we show that the RecA protein can catalyze a four-strand exchange. Specifically, it can recombine short hairpin substrates with homologous stems provided that one of the hairpins possesses a chimeric DNA/RNA backbone. This four-strand exchange reaction goes to completion in the presence of ATPgammaS and releases a stable heteroduplex upon removal of the RecA protein. Under identical conditions, strand exchange between two DNA hairpins is incomplete and generates a nascent heteroduplex that rapidly dissociates when the RecA protein is denatured. Since presynaptic filament formation does not appear to melt either type of hairpin, we propose that exchange occurs between homologously aligned duplexes that are extended and unwound within a RecA filament. The first reaction provides a mechanism for gene targeting by chimeric double-hairpin oligonucleotides while the second reaction explains the ability of the RecA protein to transiently align double-stranded DNA molecules.  相似文献   

16.
Abstract

We have characterised complexes between RecA and single-stranded homopolynucleotides by linear dichroism spectroscopy and small angle neutron scattering to investigate base pairing possibilities among DNA strands bound in a RecA filament. We find that in the presence of the non-hydrolysable cofactor ATPγS, and very likely also in the presence of ATP, a RecA fiber has three distinct DNA binding sites, each of which can bind one strand of DNA at a stoichiometry of three nucleotides per RecA monomer. The structural and hydrodynamic properties of the complexes are found to depend on the number of strands bound and on sequence complementarity among the strands. For example, RecA-[homopolymer]3-ATPγS complexes aggregate when either of the strands bound in sites I and II is complementary to the strand bound in site III. We have also studied the RecA catalysed annealing of complementary homopolymers and find it to be most efficient when two strands of one homopolymer are bound per RecA filament prior to the addition of the complementary homopolymer. These results suggest that a DNA strand bound in site III can base-pair with either of the strands in sites I and II, whereas the latter strands are unable to base-pair with each other.  相似文献   

17.
Damage to cellular DNA or interruption of chromosomal DNA synthesis leads to induction of the SOS functions in E. coli. The immediate agent of induction is the RecA protein, which proteolytically cleaves and inactivates repressors, leading to induction of genes they control. RecA protein modified by tif mutations allows expression of SOS functions in the absence of inducing treatments. We show here that tif-mutant RecA protein is more efficient than wild-type RecA protein in interacting with DNA and nucleoside triphosphate. This result suggests that formation of a complex with DNA and nucleoside triphosphate is the critical event that activates RecA protein to destroy repressors after SOS-inducing treatments, and that damage to cellular DNA promotes this reaction by providing single-stranded DNA or an active nucleoside triphosphate or both. Since dATP is the most effective nucleoside triphosphate in promoting repressor cleavage, we suggest that it is the natural cofactor of RecA protein in vivo.  相似文献   

18.
Corrinoids are essential cofactors of enzymes involved in the C1 metabolism of anaerobes. The active, super‐reduced [CoI] state of the corrinoid cofactor is highly sensitive to autoxidation. In O‐demethylases, the oxidation to inactive [CoII] is reversed by an ATP‐dependent electron transfer catalyzed by the activating enzyme (AE). The redox potential changes of the corrinoid cofactor, which occur during this reaction, were studied by potentiometric titration coupled to UV/visible spectroscopy. By applying europium(II)–diethylenetriaminepentaacetic acid (DTPA) as a reductant, we were able to determine the midpoint potential of the [CoII]/[CoI] couple of the protein‐bound corrinoid cofactor in the absence and presence of AE and/or ATP. The data revealed that the transfer of electrons from a physiological donor to the corrinoid as the electron‐accepting site is achieved by increasing the potential of the corrinoid cofactor from ?530 ± 15 mV to ?250 ± 10 mV (ESHE, pH 7.5). The first 50 to 100 mV of the shift of the redox potential seem to be caused by the interaction of nucleotide‐bound AE with the corrinoid protein or its cofactor. The remaining 150–200 mV had to be overcome by the chemical energy of ATP hydrolysis. The experiments revealed that Eu(II)–DTPA, which was already known as a powerful reducing agent, is a suitable electron donor for titration experiments of low‐potential redox centers. Furthermore, the results of this study will contribute to the understanding of thermodynamically unfavorable electron transfer processes driven by the power of ATP hydrolysis.  相似文献   

19.
Isoelectric focusing (IEF) of a somatic extract of adult worms (SEAW) yielded nine fractions. Most of the applied protein antigen was recovered in five fractions with pI values of 5.2, 4.4, 4.3, 4.0, and 3.3, respectively. The nine IEF fractions of SEAW gave a total of 37 bands following electrophoresis on polyacrylamide gel (PAGE). IEF fractionation of a somatic extract of microfilariae (SEM) yielded nine fractions. Two fractions with pI values of 4.4 and 3.2, respectively, contained most of the applied protein. The nine IEF fractions of SEM gave a total of 32 bands following PAGE. Crossed immunoelectrophoresis (CI) of SEM and SEAW against serum from hyperinfected hamsters yielded five and eight precipitation peaks, respectively. CI of SEM and SEAW against their homologous rabbit antisera gave 7 and 10 major precipitation peaks, respectively. Immunodiffusion of the nine IEF fractions from both SEM and SEAW against both their homologous and heterologous rabbit antisera indicated three and four precipitin bands peculiar only to SEM and SEAW, respectively. The remaining 23 bands from both preparations showed lines of identity.  相似文献   

20.
Two mutant Escherichia coli RecA proteins were prepared in which the ATP active site residue, Ser240, was replaced with asparagine and lysine (these amino acids are found in the corresponding positions in other bacterial RecA proteins). The S240N mutation had no discernible effect on the ATP-dependent activities of the RecA protein, indicating that serine and asparagine are functionally interchangeable at position 240. The S240K mutation, in contrast, essentially eliminated the ability of the RecA protein to utilize ATP as a nucleotide cofactor. The [S240K]RecA protein was able to catalyze the hydrolysis of dATP, however, suggesting that the absence of the 2'-hydroxyl group reduced an inhibitory interaction with the Lys240 side chain. Interestingly, the [S240K]RecA protein was able to promote an efficient LexA cleavage reaction but exhibited no strand exchange activity when dATP was provided as the nucleotide cofactor. This apparent separation of function may be attributable to the elevated S(0.5) value for dATP for the [S240K]RecA protein (490 μM, compared to 20-30 μM for the wild type and [S240N]RecA proteins), and may reflect a differential dependence of the LexA co-protease and DNA strand exchange activities on the nucleotide cofactor-mediated stabilization of the functionally-active state of the RecA-ssDNA complex.  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