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 46 毫秒
1.
In the yeast Saccharomyces cerevisiae, mismatch repair (MMR) is initiated by the binding of heterodimeric MutS homolog (MSH) complexes to mismatches that include single nucleotide and loop insertion/deletion mispairs. In in vitro experiments, the mismatch binding specificity of the MSH2-MSH6 heterodimer is eliminated if ATP is present. However, addition of the MutL homolog complex MLH1-PMS1 to binding reactions containing MSH2-MSH6, ATP, and mismatched substrate results in the formation of a stable ternary complex. The stability of this complex suggests that it represents an intermediate in MMR that is subsequently acted upon by other MMR factors. In support of this idea, we found that the replication processivity factor proliferating cell nuclear antigen (PCNA), which plays a critical role in MMR at step(s) prior to DNA resynthesis, disrupted preformed ternary complexes. These observations, in conjunction with experiments performed with streptavidin end-blocked mismatch substrates, suggested that PCNA interacts with an MSH-MLH complex formed on DNA mispairs.  相似文献   

2.
In yeast, MSH2 interacts with MSH6 to repair base pair mismatches and single nucleotide insertion/deletion mismatches and with MSH3 to recognize small loop insertion/deletion mismatches. We identified a msh6 mutation (msh6-F337A) that when overexpressed in wild type strains conferred a defect in both MSH2-MSH6- and MSH2-MSH3-dependent mismatch repair pathways. Genetic analysis suggested that this phenotype was due to msh6-F337A sequestering MSH2 and preventing it from interacting with MSH3 and MSH6. In UV cross-linking, filter binding, and gel retardation assays, the MSH2-msh6-F337A complex displayed a mismatch recognition defect. These observations, in conjunction with ATPase and dissociation rate analysis, suggested that MSH2-msh6-F337A formed an unproductive complex that was unable to stably bind to mismatch DNA.  相似文献   

3.
The interaction of the Saccharomyces cerevisiae MSH2-MSH6 complex with mispaired bases was analyzed using gel mobility shift assays and surface plasmon resonance methods. Under equilibrium binding conditions, MSH2-MSH6 bound to homoduplex DNA with a K(d) of 3.9 nM and bound oligonucleotide duplexes containing T:G, +1, +2, +4, and +10 insertion/deletion loop (IDL) mispairs with K(d) values of 0.20, 0.25, 11, 3.2, and 0.55 nM, respectively. Competition binding experiments using 65 different substrates revealed a 10-fold range in mispair discrimination. In general, base-base mispairs and a +1 insertion/deletion mispair were recognized better than intermediate sized insertion/deletion mispairs of 2-8 bases. Larger IDL mispairs (>8 bases) were recognized almost as well as the +1 IDL mispair. Recognition of mispairs by MSH2-MSH6 was influenced by sequence context, with the 6-nucleotide region surrounding the mispair being primarily responsible for influencing mispair recognition. Effects of sequences as far away as 15 nucleotides were also observed. Differential effects of ATP on the stability of MSH2-MSH6-mispair complexes suggested that base-base mispairs and the smaller IDL mispairs were recognized by a different binding mode than larger IDL mispairs, consistent with genetic experiments indicating that MSH2-MSH6 functions primarily in the repair of base-base and small IDL mispairs.  相似文献   

4.
The yeast MSH2-MSH6 complex is required to repair both base-pair and single base insertion/deletion mismatches. MSH2-MSH6 binds to mismatch substrates and displays an ATPase activity that is modulated by mispairs that are repaired in vivo. To understand early steps in mismatch repair, we analyzed mismatch repair (MMR) defective MSH2-msh6-F337A and MSH2-msh6-340 complexes that contained amino acid substitutions in the MSH6 mismatch recognition domain. While both heterodimers were defective in forming stable complexes with mismatch substrates, only MSH2-msh6-340 bound to homoduplex DNA with an affinity that was similar to that observed for MSH2-MSH6. Additional analyses suggested that stable binding to a mispair is not sufficient to initiate recruitment of downstream repair factors. Previously, we observed that MSH2-MSH6 forms a stable complex with a palindromic insertion mismatch that escapes correction by MMR in vivo. Here we show that this binding is not accompanied by either a modulation in MSH2-MSH6 ATPase activity or an ATP-dependent recruitment of the MLH1-PMS1 complex. Together, these observations suggest that early stages in MMR can be divided into distinct recognition, stable binding, and downstream factor recruitment steps.  相似文献   

5.
In eukaryotes the MSH2-MSH3 and MSH2-MSH6 heterodimers initiate mismatch repair (MMR) by recognizing and binding to DNA mismatches. The MLH1-PMS1 heterodimer then interacts with the MSH proteins at or near the mismatch site and is thought to act as a mediator to recruit downstream repair proteins. Here we analyzed five msh2 mutants that are functional in removing 3' non-homologous tails during double-strand break repair but are completely defective in MMR. Because non-homologous tail removal does not require MSH6, MLH1, or PMS1 functions, a characterization of the msh2 separation of function alleles should provide insights into early steps in MMR. Using the Taq MutS crystal structure as a model, three of the msh2 mutations, msh2-S561P, msh2-K564E, msh2-G566D, were found to map to a domain in MutS involved in stabilizing mismatch binding. Gel mobility shift and DNase I footprinting assays showed that two of these mutations conferred strong defects on MSH2-MSH6 mismatch binding. The other two mutations, msh2-S656P and msh2-R730W, mapped to the ATPase domain. DNase I footprinting, ATP hydrolysis, ATP binding, and MLH1-PMS1 interaction assays indicated that the msh2-S656P mutation caused defects in ATP-dependent dissociation of MSH2-MSH6 from mismatch DNA and in interactions between MSH2-MSH6 and MLH1-PMS1. In contrast, the msh2-R730W mutation disrupted MSH2-MSH6 ATPase activity but did not strongly affect ATP binding or interactions with MLH1-PMS1. These results support a model in which MMR can be dissected into discrete steps: stable mismatch binding and sensing, MLH1-PMS1 recruitment, and recycling of MMR components.  相似文献   

6.
The pathways whereby Sox2 scans DNA to locate its specific binding site are investigated by NMR in specific and nonspecific Sox2·DNA complexes and in a specific ternary complex with Oct1 on the Hoxb1 regulatory element. Direct transfer of Sox2 between nonspecific sites on different DNA molecules occurs without dissociation into free solution at a rate of ~10(6) M(-1) s(-1), whereas one-dimensional sliding proceeds with a diffusion constant of ≥0.1 μm(2)·s(-1). Translocation of Sox2 from one specific DNA site to another occurs via jumping, involving complete dissociation into free solution (k(d) ~5-6 s(-1)) followed by reassociation (k(a) ~5 × 10(8) M(-1) s(-1)). In the presence of Oct1 bound to an adjacent specific site, k(d) is reduced by more than 10-fold. Paramagnetic relaxation measurements, however, demonstrate that sparsely populated (<1%), transient states involving nonspecifically bound Sox2 in rapid exchange with specifically bound Sox2 are sampled in both binary Sox2·DNA- and ternary Oct1·Sox2·Hoxb1-DNA-specific complexes. Moreover, Sox2 modulates the mechanism of translocation of Oct1. Both Sox2 and the Oct1 POU(HD) domain are transiently released from the specific ternary complex by sliding to an adjacent nonspecific site, followed by direct transfer to another DNA molecule, whereas the Oct1 POU(S) domain is fixed to its specific site through direct interactions with Sox2. Intermolecular translocation of POU(HD) results in the formation of a bridged intermediate spanning two DNA molecules, enhancing the probability of complete intermolecular translocation of Oct1. By way of contrast, in the specific Oct1·DNA binary complex, POU(S) undergoes direct intermolecular translocation, whereas POU(HD) scans the DNA by sliding.  相似文献   

7.
The heterodimeric human MSH2-MSH6 protein initiates DNA mismatch repair (MMR) by recognizing mismatched bases that result from replication errors. Msh2(G674A) or Msh6(T1217D) mice that have mutations in or near the ATP binding site of MSH2 or ATP hydrolysis catalytic site of MSH6 develop cancer and have a reduced lifespan due to loss of the MMR pathway (Lin, D. P., Wang, Y., Scherer, S. J., Clark, A. B., Yang, K., Avdievich, E., Jin, B., Werling, U., Parris, T., Kurihara, N., Umar, A., Kucherlapati, R., Lipkin, M., Kunkel, T. A., and Edelmann, W. (2004) Cancer Res. 64, 517-522; Yang, G., Scherer, S. J., Shell, S. S., Yang, K., Kim, M., Lipkin, M., Kucherlapati, R., Kolodner, R. D., and Edelmann, W. (2004) Cancer Cell 6, 139-150). Mouse embryonic fibroblasts from these mice retain an apoptotic response to DNA damage. Mutant human MutSα proteins MSH2(G674A)-MSH6(wt) and MSH2(wt)-MSH6(T1219D) are profiled in a variety of functional assays and as expected fail to support MMR in vitro, although they retain mismatch recognition activity. Kinetic analyses of DNA binding and ATPase activities and examination of the excision step of MMR reveal that the two mutants differ in their underlying molecular defects. MSH2(wt)-MSH6(T1219D) fails to couple nucleotide binding and mismatch recognition, whereas MSH2(G674A)-MSH6(wt) has a partial defect in nucleotide binding. Nevertheless, both mutant proteins remain bound to the mismatch and fail to promote efficient excision thereby inhibiting MMR in vitro in a dominant manner. Implications of these findings for MMR and DNA damage signaling by MMR proteins are discussed.  相似文献   

8.
L Galio  C Bouquet    P Brooks 《Nucleic acids research》1999,27(11):2325-2331
Functional interactions of Escherichia coli MutS and MutL in mismatch repair are dependent on ATP. In this study, we show that MutS and MutL associate with immobilised DNA in a manner dependent on ATP hydrolysis and with an ATP concentration near the solution K m of the ATPase of MutS. After removal of MutS, MutL and ATP, much of the protein in this ternary complex is not stably associated, with MutL leaving the complex more rapidly than MutS. The rapid dissociation reveals a dynamic interaction with concurrent rapid association and dissociation of proteins from the DNA. Analysis by surface plasmon resonance showed that the DNA interacting with dynamically bound protein was more resistant to nuclease digestion than the DNA in MutS-DNA complexes. Non-hydrolysable analogs of ATP inhibit the formation of this dynamic complex, but permit formation of a second type of ternary complex with MutS and MutL stably bound to the immobilised DNA.  相似文献   

9.
Postreplication DNA mismatch repair is initiated by the eukaryotic protein MSH2-MSH6 or the prokaryotic protein MutS, both showing overall conserved structure and functionality. Crystal structures of MSH2-MSH6 and MutS bound to the mismatch DNA reveal a closed architecture of the clamp and the lever domains exhibiting strong contacts with the bent DNA backbone. Long molecular dynamics simulations of the human MSH2-MSH6 protein in the absence of a DNA show an altered conformation of the protein that reflects the protein's state before binding to DNA. The clamp and the lever domains of both MSH6 and MSH2 open in an asymmetric and dramatic fashion. The opening of the clamp and the lever domains in the absence of DNA is coupled to changes in the ATPase domains, which explains the experimentally observed diminished ATPase activity in DNA-free MSH2-MSH6 and illustrates the allosteric coupling between DNA binding and ATPase activity.  相似文献   

10.
During mismatch repair ATP binding and hydrolysis activities by the MutS family proteins are important for both mismatch recognition and for transducing mismatch recognition signals to downstream repair factors. Despite intensive efforts, a MutS.ATP.DNA complex has eluded crystallographic analysis. Searching for ATP analogs that strongly bound to Thermus aquaticus (Taq) MutS, we found that ADP.beryllium fluoride (ABF), acted as a strong inhibitor of several MutS family ATPases. Furthermore, ABF promoted the formation of a ternary complex containing the Saccharomyces cerevisiae MSH2.MSH6 and MLH1.PMS1 proteins bound to mismatch DNA but did not promote dissociation of MSH2.MSH6 from mismatch DNA. Crystallographic analysis of the Taq MutS.DNA.ABF complex indicated that although this complex was very similar to that of MutS.DNA.ADP, both ADP.Mg(2+) moieties in the MutS. DNA.ADP structure were replaced by ABF. Furthermore, a disordered region near the ATP-binding pocket in the MutS B subunit became traceable, whereas the equivalent region in the A subunit that interacts with the mismatched nucleotide remained disordered. Finally, the DNA binding domains of MutS together with the mismatched DNA were shifted upon binding of ABF. We hypothesize that the presence of ABF is communicated between the two MutS subunits through the contact between the ordered loop and Domain III in addition to the intra-subunit helical lever arm that links the ATPase and DNA binding domains.  相似文献   

11.
The DNA-dependent protein kinase (DNA-PK), consisting of Ku and the DNA-PK catalytic subunit (DNA-PKcs), and the DNA ligase IV-XRCC4 complex function together in the repair of DNA double-strand breaks by non-homologous end joining. These protein complexes are also required for the completion of V(D)J recombination events in immune cells. Here we demonstrate that the DNA ligase IV-XRCC4 complex binds specifically to the ends of duplex DNA molecules and can act as a bridging factor, linking together duplex DNA molecules with complementary but non-ligatable ends. Although the DNA end-binding protein Ku inhibited DNA joining by DNA ligase IV-XRCC4, it did not prevent this complex from binding to DNA. Instead, DNA ligase IV-XRCC4 and Ku bound simultaneously to the ends of duplex DNA molecules. DNA ligase IV-XRCC4 and DNA-PKcs also formed complexes at the ends of DNA molecules, but DNA-PKcs did not inhibit ligation. Interestingly, DNA-PKcs stimulated intermolecular ligation by DNA ligase IV-XRCC4. In the presence of DNA-PK, the majority of the joining events catalyzed by DNA ligase IV-XRCC4 were intermolecular because Ku inhibited intramolecular ligation, but DNA-PKcs still stimulated intramolecular ligation. We suggest that DNA-PKcs-containing complexes formed at DNA ends enhance the association of DNA ends via protein-protein interactions, thereby stimulating intermolecular ligation.  相似文献   

12.
Proliferating cell nuclear antigen (PCNA) is thought to play a role in DNA mismatch repair at the DNA synthesis step as well as in an earlier step. Studies showing that PCNA interacts with mispair-binding protein complexes, MSH2.MSH3 and MSH2.MSH6, and that PCNA enhances MSH2.MSH6 mispair binding specificity suggest PCNA may be involved in mispair recognition. Here we show that PCNA and MSH2.MSH6 form a stable ternary complex with a homoduplex (G/C) DNA, but MSH2.MSH6 binding to a heteroduplex (G/T) DNA disrupts MSH2.MSH6 binding to PCNA. We also found that the addition of ATP or adenosine 5'-O-(thiotriphosphate) restores MSH2.MSH6 binding to PCNA, presumably by disrupting MSH2.MSH6 binding to the heteroduplex (G/T) DNA. These results support a model in which MSH2.MSH6 binds to PCNA loaded on newly replicated DNA and is transferred from PCNA to mispaired bases in DNA.  相似文献   

13.
The ability of DNA gyrase (Gyr) to wrap the DNA strand around itself allows Gyr to introduce negative supercoils into DNA molecules. It has been demonstrated that the deletion of the C-terminal DNA-binding domain of the GyrA subunit abolishes the ability of Gyr to wrap the DNA strand and catalyze the supercoiling reaction (Kampranis, S. C., and Maxwell, A. (1996) Proc. Natl. Acad. Sci. U. S. A. 93, 14416-14421). By using this mutant Gyr, Gyr (A59), we have studied effects of Gyr-mediated wrapping of the DNA strand on its replicative function and its interaction with the quinolone antibacterial drugs. We find that Gyr (A59) can support oriC DNA replication in vitro. However, Gyr (A59)-catalyzed decatenation activity is not efficient enough to complete the decatenation of replicating daughter DNA molecules. As is the case with topoisomerase IV, the active cleavage and reunion activity of Gyr is required for the formation of the ternary complex that can arrest replication fork progression in vitro. Although the quinolone drugs stimulate the covalent Gyr (A59)-DNA complex formation, the Gyr (A59)-quinolone-DNA ternary complexes do not arrest the progression of replication forks. Thus, the quinolone-induced covalent topoisomerase-DNA complex formation is necessary but not sufficient to cause the inhibition of DNA replication. We also assess the stability of ternary complexes formed with Gyr (A59), the wild type Gyr, or topoisomerase IV. The ternary complexes formed with Gyr (A59) are more sensitive to salt than those formed with either the wild type Gyr or topoisomerase IV. Furthermore, a competition experiment demonstrates that the ternary complexes formed with Gyr (A59) readily disassociate from the DNA, whereas the ternary complexes formed with either the wild type Gyr or topoisomerase IV remain stably bound. Thus, Gyr-mediated wrapping of the DNA strand is required for the formation of the stable Gyr-quinolone-DNA ternary complex that can arrest replication fork progression.  相似文献   

14.
Mammalian MutL homologues function in DNA mismatch repair (MMR) after replication errors and in meiotic recombination. Both functions are initiated by a heterodimer of MutS homologues specific to either MMR (MSH2-MSH3 or MSH2-MSH6) or crossing over (MSH4-MSH5). Mutations of three of the four MutL homologues (Mlh1, Mlh3, and Pms2) result in meiotic defects. We show herein that two distinct complexes involving MLH3 are formed during murine meiosis. The first is a stable association between MLH3 and MLH1 and is involved in promoting crossing over in conjunction with MSH4-MSH5. The second complex involves MLH3 together with MSH2-MSH3 and localizes to repetitive sequences at centromeres and the Y chromosome. This complex is up-regulated in Pms2-/- males, but not females, providing an explanation for the sexual dimorphism seen in Pms2-/- mice. The association of MLH3 with repetitive DNA sequences is coincident with MSH2-MSH3 and is decreased in Msh2-/- and Msh3-/- mice, suggesting a novel role for the MMR family in the maintenance of repeat unit integrity during mammalian meiosis.  相似文献   

15.
Formation of a ternary complex between human MutSalpha, MutLalpha, and heteroduplex DNA has been demonstrated by surface plasmon resonance spectroscopy and electrophoretic gel shift methods. Formation of the hMutLalpha.hMutSalpha.heteroduplex complex requires a mismatch and ATP hydrolysis, and depends on DNA chain length. Ternary complex formation was supported by a 200-base pair G-T heteroduplex, a 100-base pair substrate was somewhat less effective, and a 41-base pair heteroduplex was inactive. As judged by surface plasmon resonance spectroscopy, ternary complexes produced with the 200-base pair G-T DNA contained approximately 0.8 mol of hMutLalpha/mol of heteroduplex-bound hMutSalpha. Although the steady-state levels of the hMutLalpha.hMutSalpha. heteroduplex were substantial, this complex was found to turn over, as judged by surface plasmon resonance spectroscopy and electrophoretic gel shift analysis. With the former method, the majority of the complexes dissociated rapidly upon termination of protein flow, and dissociation occurred in the latter case upon challenge with competitor DNA. However, ternary complex dissociation as monitored by gel shift assay was prevented if both ends of the heteroduplex were physically blocked with streptavidin.biotin complexes. This observation suggests that, like hMutSalpha, the hMutLalpha.hMutSalpha complex can migrate along the helix contour to dissociate at DNA ends.  相似文献   

16.
The binding of two chimeric proteins, consisting of the N-terminal or C-terminal DNA binding domain of Tn916 Int fused to maltose binding protein, to specific oligonucleotide substrates was analyzed by gel mobility shift assay. The chimeric protein with the N-terminal domain formed two complexes of different electrophoretic mobilities. The faster-moving complex, whose formation displayed no cooperativity, contained two protein monomers bound to a single DNA molecule. The slower-moving complex, whose formation involved cooperative binding (Hill coefficient > 1.0), contained four protein monomers bound to a single DNA molecule. Methylation interference experiments coupled with the analysis of protein binding to mutant oligonucleotide substrates showed that formation of the faster-moving complex containing two protein monomers required the presence of two 11-bp direct repeats (called DR2) in direct orientation. Formation of the slower-moving complex required only a single DR2 repeat. Binding of the N-terminal domains in vivo could serve to position two Int monomers on the DNA near each end of the transposon and assist in bringing together the ends of the transposon so that excision can occur. The chimeric protein with the C-terminal domain of Int also formed two complexes of different electrophoretic mobilities. The major, slower-moving complex, whose formation involved cooperative binding, contained two protein molecules bound to one DNA molecule. This finding suggested that while the C-terminal domain of Int can bind DNA as a monomer, a cooperative interaction between two monomers of the C-terminal domain may help to bring the ends of the transposon together during excision.  相似文献   

17.
In eukaryotic mismatch repair (MMR) MSH2-MSH6 initiates the repair of base-base and small insertion/deletion mismatches while MSH2-MSH3 repairs larger insertion/deletion mismatches. Here, we show that the msh2Delta1 mutation, containing a complete deletion of the conserved mismatch recognition domain I of MSH2, conferred a separation of function phenotype with respect to MSH2-MSH3 and MSH2-MSH6 functions. Strains bearing the msh2Delta1 mutation were nearly wild-type in MSH2-MSH6-mediated MMR and in suppressing recombination between DNA sequences predicted to form mismatches recognized by MSH2-MSH6. However, these strains were completely defective in MSH2-MSH3-mediated MMR and recombination functions. This information encouraged us to analyze the contributions of domain I to the mismatch binding specificity of MSH2-MSH3 in genetic and biochemical assays. We found that domain I in MSH2 contributed a non-specific DNA binding activity while domain I of MSH3 appeared important for mismatch binding specificity and for suppressing non-specific DNA binding. These observations reveal distinct requirements for the MSH2 DNA binding domain I in the repair of DNA mismatches and suggest that the binding of MSH2-MSH3 to mismatch DNA involves protein-DNA contacts that appear very different from those required for MSH2-MSH6 mismatch binding.  相似文献   

18.
Oxidation of G in DNA yields 8-oxo-G (GO), a mutagenic lesion that leads to misincorporation of A opposite GO. In E. coli, GO in GO:C base pairs is removed by MutM, and A in GO:A mispairs is removed by MutY. In S. cerevisiae, mutations in MSH2 or MSH6 caused a synergistic increase in mutation rate in combination with mutations in OGG1, which encodes a MutM homolog, resulting in a 140- to 218-fold increase in the G:C-to-T:A transversion rate. Consistent with this, MSH2-MSH6 complex bound to GO:A mispairs and GO:C base pairs with high affinity and specificity. These data indicate that in S. cerevisiae, MSH2-MSH6-dependent mismatch repair is the major mechanism by which misincorporation of A opposite GO is corrected.  相似文献   

19.
MuB, a protein essential for replicative DNA transposition by the bacteriophage Mu, is an ATPase that assembles into a polymeric complex on DNA. We used total internal reflection fluorescence microscopy to observe the behavior of MuB polymers on single molecules of DNA. We demonstrate that polymer assembly is initiated by a stochastic nucleation event. After nucleation, polymer assembly occurs by a mechanism involving the sequential binding of small units of MuB. MuB that bound to A/T-rich regions of the DNA assembled into large polymeric complexes. In contrast, MuB that bound outside of the A/T-rich regions failed to assemble into large oligomeric complexes. Our data also show that MuB does not catalyze multiple rounds of ATP hydrolysis while remaining bound to DNA. Rather, a single ATP is hydrolyzed, then MuB dissociates from the DNA. Finally, we show that "capping" of the enhanced green fluorescent protein-MuB polymer ends with unlabeled MuB dramatically slows, but does not halt, dissociation. This suggests that MuB dissociation occurs through both an end-dependent mechanism and a slower mechanism wherein subunits dissociate from the polymer interior.  相似文献   

20.
Quantitative studies of interactions between the Escherichia coli replication factor DnaC protein and the DnaB helicase have been performed using sedimentation velocity and fluorescence energy transfer techniques. The applied novel analysis of the sedimentation data allows us to construct thermodynamic rigorous binding isotherms without any assumption as to the relationship between the observed molecular property of the complexes formed, the average sedimentation coefficient, or the degree of binding. Experiments have been performed with the fluorescein-modified DnaB helicase, which allows an exclusive monitoring of the DnaB-DnaC complex formation. The DnaC binding to the unmodified helicase has been characterized in competition experiments. The data establish that, in the presence of the ATP analog AMP-PNP, or ADP, a maximum of six DnaC monomers bind cooperatively to the DnaB hexamer. The positive cooperative interactions are limited to the two neighboring DnaC molecules. Analyses using a statistical thermodynamic hexagon model indicate that, under the solution conditions examined, the affinity is characterized by the intrinsic binding constant K=1.4(+/-0.5)x10(5)M(-1) and cooperativity parameter sigma=21+/-5. These data suggest strongly that the DnaC-DnaB complex exists in vivo as a mixture of complexes with a different number of bound DnaC molecules, although the complex with six DnaC molecules bound dominates the distribution. The DnaC nucleotide-binding site is not involved in the stabilization of the complex. Moreover, the hydrolysis of NTP bound to the helicase or the DnaC is not required for the release of the DnaC protein from the complex. The single-stranded DNA (ssDNA) bound to the helicase does not affect the DnaC protein binding. However, in the presence of the DNA, there is a significant difference in the energetics and structure of the ternary complex, DnaC-DnaB-ssDNA, formed in the presence of AMP-PNP as compared to ADP. The topology of the ternary complex DnaC-DnaB-ssDNA has been determined using the fluorescence energy transfer method. In solution, the DnaC protein-binding site is located on the large 33 kDa domain of the DnaB helicase. The significance of the results in the functioning of the DnaB helicase-DnaC protein complex is discussed.  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