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.
Kozlov AG  Lohman TM 《Biochemistry》2006,45(16):5190-5205
We have previously shown that the linkage of temperature-dependent protonation and DNA base unstacking equilibria contribute significantly to both the negative enthalpy change (DeltaH(obs)) and the negative heat capacity change (DeltaC(p,obs)) for Escherichia coli SSB homotetramer binding to single-stranded (ss) DNA. Using isothermal titration calorimetry we have now examined DeltaH(obs) over a much wider temperature range (5-60 degrees C) and as a function of monovalent salt concentration and type for SSB binding to (dT)(70) under solution conditions that favor the fully wrapped (SSB)(65) complex (monovalent salt concentration >or=0.20 M). Over this wider temperature range we observe a strongly temperature-dependent DeltaC(p,obs). The DeltaH(obs) decreases as temperature increases from 5 to 35 degrees C (DeltaC(p,obs) <0) but then increases at higher temperatures up to 60 degrees C (DeltaC(p,obs) >0). Both salt concentration and anion type have large effects on DeltaH(obs) and DeltaC(p,obs). These observations can be explained by a model in which SSB protein can undergo a temperature- and salt-dependent conformational transition (below 35 degrees C), the midpoint of which shifts to higher temperature (above 35 degrees C) for SSB bound to ssDNA. Anions bind weakly to free SSB, with the preference Br(-) > Cl(-) > F(-), and these anions are then released upon binding ssDNA, affecting both DeltaH(obs) and DeltaC(p,obs). We conclude that the experimentally measured values of DeltaC(p,obs) for SSB binding to ssDNA cannot be explained solely on the basis of changes in accessible surface area (ASA) upon complex formation but rather result from a series of temperature-dependent equilibria (ion binding, protonation, and protein conformational changes) that are coupled to the SSB-ssDNA binding equilibrium. This is also likely true for many other protein-nucleic acid interactions.  相似文献   

2.
3.
T M Lohman  W Bujalowski 《Biochemistry》1988,27(7):2260-2265
We have examined the binding of the oligonucleotide dT (pT)34 to the Escherichia coli SSB protein as a function of NaCl and MgCl2 concentration (25 degrees C, pH 8.1) by monitoring the quenching of the intrinsic protein fluorescence. We find two binding sites for dT(pT)34 per single strand binding (SSB) protein tetramer, with each site possessing widely different affinities depending on the salt concentration. At 200 mM NaCl, we observe nearly stoichiometric binding of dT(pT)34 to both binding sites within the SSB tetramer, although a difference in the affinities is still apparent. However, when the NaCl concentration is lowered, the overall affinity of dT(pT)34 for the second site on the SSB tetramer decreases dramatically. At 1.5 mM NaCl, only a single molecule of dT(pT)34 can bind per SSB tetramer, even with a 10-fold molar excess of dT(pT)34. MgCl2 is effective at 100-fold lower concentrations than NaCl in promoting the binding of the second molecule of dT(pT)34. This binding behavior reflects an intrinsic property of the SSb tetramer, since it is also observed upon binding of smaller oligonucleotides, and the simplest explanation is that a salt-dependent negative cooperativity exists between DNA binding sites within the SSB tetramer. This phenomenon is also responsible for the transition between the two SSB-single strand (ss) polynucleotide binding modes that cover 35 and 56 nucleotides per tetramer [Bujalowski, W., & Lohman, T. M. (1986) Biochemistry 25, 7799-7802]. Extreme negative cooperativity stabilizes the (SSB)35 binding mode, in which the SSB tetramer binds tightly to ss DNA with only two of its subunits while the other two subunits remain unligated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
R R?mer  U Schomburg  G Krauss  G Maass 《Biochemistry》1984,23(25):6132-6137
The interaction of the Escherichia coli single-stranded DNA binding protein (SSB) with oligo- and poly-nucleotides has been studied by 270-MHz 1H NMR spectroscopy and fast kinetic techniques. d(pT)8 and poly(dT) were used to study noncooperative and cooperative binding, respectively. The H6, H1', and CH3 resonances of d(pT)8 are high-field shifted by less than 0.05 ppm, and H8 and H2 of poly(dA) are low-field shifted upon complexation. The protein resonances remain virtually unshifted. The small shifts upon complexation provide no evidence for extensive stacking interactions between the nucleotide bases and aromatic amino acid side chains of SSB. The d(pT)8 and poly(dA) signals are broadened to about 30 Hz whereas the resonances of poly(dT) are broadened beyond detection upon stoichiometric complexation. Continuous broadening of all poly(dT) signals even at a 10-fold excess of poly(dT) indicates fast exchange of SSB between different binding sites. Dissociation and reassociation rates determined from stopped-flow experiments are too slow by at least 2 orders of magnitude to account for the experimental line widths. Therefore, we conclude that SSB translocates without dissociation from the DNA template. A model for the translocation is outlined. It is based on partial dissociation of octamer sections of poly(dT) from the complex with a rate constant as previously published for the dissociation of d(pT)8 from SSB.  相似文献   

5.
Isothermal titration calorimetry has been used to investigate the thermodynamic parameters of the binding of thymidine (dT) and ATP to herpes simplex virus type 1 thymidine kinase (HSV1 TK). Binding follows a sequential pathway in which dT binds first and ATP second. The free enzyme does not bind ATP, whose binding site becomes only accessible in the HSV1 TK.dT complex. At pH 7.5 and 25 degrees C, the binding constants are 1.9 x 10(5) m(-1) for dT and 3.9 x 10(6) m(-1) for ATP binding to the binary HSV1 TK.dT complex. Binding of both substrates is enthalpy-driven and opposed by a large negative entropy change. The heat capacity change (DeltaCp) obtained from DeltaH in the range of 10-25 degrees C is -360 cal K(-1) mol(-1) for dT binding and -140 cal K(-1) mol(-1) for ATP binding. These large DeltaCp values are incompatible with a rigid body binding model in which the dT and ATP binding sites pre-exist in the free enzyme. Values of DeltaCp and TDeltaS strongly indicate large scale conformational adaptation of the active site in sequential substrate binding. The conformational changes seem to be more pronounced in dT binding than in the subsequent ATP binding. Considering the crystal structure of the ternary HSV1 TK.dT.ATP complex, a large movement in the dT binding domain and a smaller but substantial movement in the LID domain are proposed to take place when the enzyme changes from the substrate-free, presumably more open and less ordered conformation to the closed and compact conformation of the ternary enzyme-substrate complex.  相似文献   

6.
A technique based on fluorescence polarization (anisotropy) was used to measure the binding of antibodies to DNA under a variety of conditions. Fluorescein-labeled duplexes of 20 bp in length were employed as the standard because they are stable even at low ionic strength yet sufficiently short so that both arms of an IgG cannot bind to the same duplex. IgG Jel 274 binds duplexes in preference to single-stranded DNA; in 80 mM NaCl Kobs for (dG)20.(dC)20 is 4.1x10(7) M-1 compared with 6.4x10(5) M-1 for d(A5C10A5). There is little sequence specificity, but the interaction is very dependent on ionic strength. From plots of log Kobs against log[Na+] it was deduced that five or six ion pairs are involved in complex formation. At low ionic strength,Kobs is independent of temperature and complex formation is entropy driven with DeltaH degrees obs and DeltaC degrees p,obs both zero. In contrast, in 80 mM NaCl DeltaC degrees p,obs is -630 and -580 cal mol-1K-1 for [d(TG)]10.[d(CA)]10 and (dG)20.(dC)20 respectively. IgG Jel 241 also binds more tightly to duplexes than single-stranded DNA, but sequence preferences were apparent. The values for Kobs to [d(AT)]20 and [d(GC)]20 are 2.7x10(8) and 1.3x10(8) M-1 respectively compared with 5.7x10(6) M-1 for both (dA)20. (dT)20 and (dG)20.(dC)20. As with Jel 274, the binding of Jel 241 is very dependent on ionic strength and four or five ionic bonds are involved in complex formation with all the duplex DNAs which were tested. DeltaC degrees p,obs for Jel 241 binding to [d(AT)]20 was negative (-87 cal mol-1K-1) in 80 mM NaCl but was zero at high ionic strength (130 mM NaCl). Therefore, for duplex-specific DNA binding antibodies DeltaC degrees p,obs is dependent on [Na+] and a large negative value does not correlate with sequence-specific interactions.  相似文献   

7.
The Km and vmax values for oligothymidylates d(pT)2-16 in reaction of 3'-5'-exonuclease hydrolysis catalyzed by Klenow fragment were measured in the absence and presence of poly(dA) template without the poly(dA), the Km values for oligonucleotides are slightly dependent on their length. The rate of oligothymidylates hydrolysis increases with their length and for d(pT)16 it is about 190-times higher than for d(pT)2. The addition on poly(dA) does not lead to an essential change of the Km values for d(pT)2-16, but raises the rate of d(pT)2-7 hydrolysis 2-17-fold and at the same time lowers the efficiency of d(pT)8-16 hydrolysis. The Km values for d(pC)10, d(pA)19 and d(pT)10 are nearly the same. However the velocity of d(pC)10 hydrolysis is approximately 1,2 and 7,8-times higher than for d(pA)10 and d(pC)10, respectively d(pC)10, d(pA)10 and d(pT)10 under conditions of interaction with the template-binding site raise the rate of hydrolysis of d(pT)2 combined with the exonuclease center, with various efficiency. Under similar conditions, d(pT)8, d(pT)10 and d(pT)16 as templates activated hydrolysis of d(pT)2. The dependence of the Klenow fragment exonuclease activity both on the length and structure of the template and on the length of the hydrolyzed oligonucleotide was suggested.  相似文献   

8.
Kozlov AG  Lohman TM 《Biochemistry》2002,41(39):11611-11627
The kinetic mechanism of transfer of the homotetrameric Escherichia coli SSB protein between ssDNA molecules was studied using stopped-flow experiments. Dissociation of SSB from the donor ssDNA was monitored after addition of a large excess of unlabeled acceptor ssDNA by using either SSB tryptophan fluorescence or the fluorescence of a ssDNA labeled with an extrinsic fluorophore [fluorescein (F) or Cy3]. The dominant pathway for SSB dissociation occurs by a "direct transfer" mechanism in which an intermediate composed of two DNA molecules bound to one SSB tetramer forms transiently prior to the release of the acceptor DNA. When an initial 1:1 SSB-ssDNA complex is formed with (dT)(70) in the fully wrapped (SSB)(65) mode so that all four SSB subunits are bound to (dT)(70), the formation of the ternary intermediate complex occurs slowly with an apparent bimolecular rate constant, k(2,app), ranging from 1.2 x 10(3) M(-1) s(-1) (0.2 M NaCl) to approximately 5.1 x 10(3) M(-1) s(-1) (0.4 M NaBr), and this rate limits the overall rate of the transfer reaction (pH 8.1, 25 degrees C). These rate constants are approximately 7 x 10(5)- and approximately 7 x 10(4)-fold lower, respectively, than those measured for binding of the same ssDNA to an unligated SSB tetramer to form a singly ligated complex. However, when an initial SSB-ssDNA complex is formed with (dT)(35) so that only two SSB subunits interact with the DNA in an (SSB)(35) complex, the formation of the ternary intermediate occurs much faster with a k(2,app) ranging from >6.3 x 10(7) M(-1) s(-1) (0.2 M NaCl) to 2.6 x 10(7) M(-1) s(-1) (0.4 M NaBr). For these experiments, the rate of dissociation of the donor ssDNA determines the overall rate of the transfer reaction. Hence, an SSB tetramer can be transferred from one ssDNA molecule to another without proceeding through a free protein intermediate, and the rate of transfer is determined by the availability of free DNA binding sites within the initial SSB-ssDNA donor complex. Such a mechanism may be used to recycle SSB tetramers between old and newly formed ssDNA regions during lagging strand DNA replication.  相似文献   

9.
This work describes differential effects of solvent in complexes of the aminoglycoside phosphotransferase(3')-IIIa (APH) with different aminoglycosides and the detection of change in solvent structure at specific sites away from substrates. Binding of kanamycins to APH occurs with a larger negative DeltaH in H2O relative to D2O (DeltaDeltaH(H2O-D2O) < 0), while the reverse is true for neomycins. Unusually large negative DeltaCp values were observed for binding of aminoglycosides to APH. DeltaCp for the APH-neomycin complex was -1.6 kcal x mol(-1) x deg(-1). A break at 30 degrees C was observed in the APH-kanamycin complex yielding DeltaCp values of -0.7 kcal x mol(-1) x deg(-1) and -3.8 kcal x mol(-1) x deg(-1) below and above 30 degrees C, respectively. Neither the change in accessible surface area (DeltaASA) nor contributions from heats of ionization were sufficient to explain the large negative DeltaCp values. Most significantly, 15N-1H HSQC experiments showed that temperature-dependent shifts of the backbone amide protons of Leu 88, Ser 91, Cys 98, and Leu143 revealed a break at 30 degrees C only in the APH-kanamycin complex in spectra collected between 21 degrees C and 38 degrees C. These amino acids represent solvent reorganization sites that experience a change in solvent structure in their immediate environment as structurally different ligands bind to the enzyme. These residues were away from the substrate binding site and distributed in three hydrophobic patches in APH. Overall, our results show that a large number of factors affect DeltaCp and binding of structurally different ligand groups cause different solvent structure in the active site as well as differentially affecting specific sites away from the ligand binding site.  相似文献   

10.
The equilibrium binding of the cytotoxic plant alkaloid berberine to various DNAs and energetics of the interaction have been studied. At low ratios of bound alkaloid to base pair, the binding exhibited cooperativity to natural DNAs having almost equal proportions of AT and GC sequences. In contrast, the binding was non-cooperative to DNAs with predominantly high AT or GC sequences. Among the synthetic DNAs, cooperative binding was observed with poly(dA).poly(dT) and poly(dG).poly(dC) while non-cooperative binding was seen with poly(dA-dT).poly(dA-dT) and poly(dG-dC).poly(dG-dC). Both cooperative and non-cooperative bindings were remarkably dependent on the salt concentration of the media. Linear plots of ln K(a) versus [Na(+)] for poly(dA).poly(dT) and poly(dA-dT).poly(dA-dT) showed the release of 0.56 and 0.75 sodium ions respectively per bound alkaloid. Isothermal titration calorimetry results revealed the binding to be exothermic and favoured by both enthalpy and entropy changes in all DNAs except the two AT polymers and AT rich DNA, where the same was predominantly entropy driven. Heat capacity values (DeltaCp(o)) of berberine binding to poly(dA).poly(dT), poly(dA-dT).poly(dA-dT), Clostridium perfringens and calf thymus DNA were -98, -140, -120 and -110 cal/mol K respectively. This study presents new insights into the binding dependent base pair heterogeneity in DNA conformation and the first complete thermodynamic profile of berberine binding to DNAs.  相似文献   

11.
We have employed a variety of physical methods to study the equilibrium melting and temperature-dependent conformational dynamics of dA.dT tracts in fractionated synthetic DNA polymers and in well-defined fragments of kinetoplast DNA (kDNA). Using circular dichroism (CD), we have detected a temperature-dependent, "premelting" event in poly(dA).poly(dT) which exhibits a midpoint near 37 degrees C. Significantly, we also detect this CD "premelting" behavior in a fragment of kDNA. By contrast, we do not observe this "premelting" behavior in the temperature-dependent CD spectra of poly[d(AT)].poly[d(AT)], poly(dG).poly(dC), poly[d(GC)].poly[d(GC)], or calf thymus DNA. Thus, poly(dA).poly(dT) and kDNA exhibit a common CD-detected "premelting" event which is absent in the other duplex systems studied in this work. Furthermore, we find that the anomalous electrophoretic retardation of the kDNA fragments we have investigated disappears at temperatures above approximately 37 degrees C. We also observe that the rotational dynamics of poly(dA).poly(dT) and kDNA as assessed by singlet depletion anisotropy decay (SDAD) and electric birefringence decay (EBD) also display a discontinuity near 37 degrees C, which is not observed for the other duplex systems studied. Thus, in the aggregate, our static and dynamic measurements suggest that the homo dA.dT sequence element [common to both poly(dA).poly(dT) and kDNA] is capable of a temperature-dependent equilibrium between at least two helical states in a temperature range well below that required to induce global melting of the host duplex. We suggest that this "preglobal" melting event may correspond to the thermally induced "disruption" of "bent" DNA.  相似文献   

12.
The comparison of the Km and Vmax values for the various primers was carried out. The primers were either completely complementary to the template or contained the non-complementary bases in different positions from the 3'-end. The number of the bases from the 3'-end to the noncomplementary nucleotide but not the primers length was supposed to determine the efficiency of the interaction of the primers containing noncomplementary bases with the enzyme. The Km values for d[(pC) (pT)7] (1.2 microM), d[(pC)3(pT)7] (2.5 microM, d[(pT)2pC(pT)7] (1.4 microM)d[(pT)4pC(pT)5(4.3 microM); d[(pT)7pC(pT)2] (11 microM) are comparable with the Km values for d(pT)7 (1.4 microM); d(pT)5 (4.2 microM) and d(pT)3 (15 mkM), respectively, but not for the decathymidilate d[(Tp)9T] (0.23 microM). The complementary interaction between the first nucleotide from the 3'-end of the primer and the template appear to play the particular role in the interaction of the enzyme with the primer. The Km values for d[(pT)10pC] and d[(pA)9pC] (with the corresponding templates) are 38 and 6 times the ones for d[(Tp)10T] and d(pA)10. However, the Km values for d[(pA)9p(rib)] (0.56 microM) which contains the deoxyribozylurea residue at the 3'-end is practically equal to the Km for d(pA)9 (0.56 microM). The Vmax values for d[(pT)10pC] and d[(pA)9pC] are 1.7 and 2.3 times the values for d[(Tp)10T] and d(pA)10, respectively. The primer affinity decreases, just as its conversion rate increases when the noncomplementary base in the primer is transferred from the 5'-to 3'-end; that results in the rate of primers elongation decrease in total.  相似文献   

13.
The poly(dA) dependent T4 polynucleotide ligase catalyzed polymerization of oligodeoxythymidylates is dependent upon duplex stability. The antibiotics ethidium bromide, netropsin and Hoechst 33258 stabilize the duplex poly(dA) . P(dT)n (n = 6-10) to thermal denaturation. Ethidium bromide to DNA ratio of 1.25 and netropsin or Hoechst 33258 to DNA ratio of 0.1 the Tm of d(pT) 10 . poly (dA) was increased by 10 degrees and 25 degrees C respectively. The T4 polynucleotide ligase activity was not inhibited under these conditions and temperature optimum of joining of d(pT) 10 . poly(dA) was increased 5 degrees to 10 degrees by the binding of the antibiotics. Duplexes containing shorter oligodeoxythymidylates required lower concentrations of the antibiotics netropsin or Hoechst 33258 to show no inhibition of T4 polynucleotide ligase. The temperature optima of joining the duplexes d(pT)6 . POLY(DA) and d(pT) 8 . poly(dA) were increased by 5 degrees C upon binding of the antibiotics. Polyacrylamide gel analysis of the T4 polynucleotide ligase catalyzed joining of the oligodeoxythymidylates showed that the presence of antibiotics affected the product distribution of the polymerized oligomers.  相似文献   

14.
The Escherichia coli wild-type single strand binding (SSB) protein is a stable tetramer that binds to single-stranded (ss) DNA in its role in DNA replication, recombination and repair. The ssb-1 mutation, a substitution of tyrosine for histidine-55 within the SSB-1 protein, destabilizes the tetramer with respect to monomers, resulting in a temperature-sensitive defect in a variety of DNA metabolic processes, including replication. Using quenching of the intrinsic SSB-1 tryptophan fluorescence, we have examined the equilibrium binding of the oligonucleotide, dT(pT)15, to the SSB-1 protein in order to determine whether a ssDNA binding site exists within individual SSB-1 monomers or whether the formation of the SSB tetramer is necessary for ssDNA binding. At high SSB-1 protein concentrations, such that the tetramer is stable, we find that four molecules of dT(pT)15 bind per tetramer in a manner similar to that observed for the wild-type SSB tetramer; i.e. negative co-operativity is observed for ssDNA binding to the SSB-1 protomers. As a consequence of this negative co-operativity, binding is biphasic, with two molecules of dT(pT)15 binding to the tetramer in each phase. However, the intrinsic binding constant, K16, for the SSB-1 protomer-dT(pT)15 interaction is a factor of 3 lower than for the wild-type protomer interaction and the negative co-operativity parameter, sigma 16, is larger in the case of the SSB-1 tetramer, indicating a lower degree of negative co-operativity. At lower SSB-1 concentrations, SSB-1 monomers bind dT(pT)15 without negative co-operativity; however, the intrinsic affinity of dT(pT)15 for the monomer is a factor of approximately 10 lower than for the protomer (50 mM-NaCl, pH 8.1, 25 degrees C). Therefore, an individual SSB-1 monomer does possess an independent ssDNA binding site; hence formation of the tetramer is not required for ssDNA binding, although tetramer formation does increase the binding affinity significantly. These data also show that the negative co-operativity among ssDNA binding sites within an SSB tetramer is an intrinsic property of the tetramer. On the basis of these studies, we discuss a modified explanation for the temperature-sensitivity of the ssb-1 phenotype.  相似文献   

15.
Ca2+,Mg2+-dependent DNAse from sea urchin embryos is specific to the secondary structure of substrates irrespective of the nature of activating cations. The enzyme does not split synthetic single-stranded oligo and polynucleotides, such as d(pTpTpTpCpC), d(pGpGpTpTpT). d(pApApTpTpC), d(pGpApApTpTpC), d(pA)5-poly(dT), d(pApApTpTpC)-poly(dT), poly(dA) and poly (dT) and hydrolyses the double-stranded substrates poly d(AT), poly (dA) . poly (dT) and highly polymerized DNA. Native double-stranded DNA from salmon and phage T7 is split by the enzyme at a higher rate than that of denaturated DNA of salmon and single-stranded DNA of phage M13. The high rate of poly(dA) . poly(dT) and poly d(AT) hydrolysis and the stability of poly(dG) . poly(dC) to the effect of the enzyme suggest a certain specificity of the enzyme to the nature of nitrogenous bases at the hydrolyzed phosphodiester bond of the substrate.  相似文献   

16.
The co-operative nature of the binding of the Escherichia coli single strand binding protein (SSB) to single-stranded nucleic acids has been examined over a range of salt concentrations (NaCl and MgCl2) to determine if different degrees of binding co-operativity are associated with the two SSB binding modes that have been identified recently. Quantitative estimates of the binding properties, including the co-operativity parameter, omega, of SSB to single-stranded DNA and RNA homopolynucleotides have been obtained from equilibrium binding isotherms, at high salt (greater than or equal to 0.2 M-NaCl), by monitoring the fluorescence quenching of the SSB upon binding. Under these high salt conditions, where only the high site size SSB binding mode exists (65 +/- 5 nucleotides per tetramer), we find only moderate co-operativity for SSB binding to both DNA and RNA, (omega = 50 +/- 10), independent of the concentration of salt. This value for omega is much lower than most previous estimates. At lower concentrations of NaCl, where the low site size SSB binding mode (33 +/- 3 nucleotides/tetramer) exists, but where SSB affinity for single-stranded DNA is too high to estimate co-operativity from classical binding isotherms, we have used an agarose gel electrophoresis technique to qualitatively examine SSB co-operativity with single-stranded (ss) M13 phage DNA. The apparent binding co-operativity increases dramatically below 0.20 M-NaCl, as judged by the extremely non-random distribution of SSB among the ssM13 DNA population at low SSB to DNA ratios. However, the highly co-operative complexes are not at equilibrium at low SSB/DNA binding densities, but are formed only transiently when SSB and ssDNA are directly mixed at low concentrations of NaCl. The conversions of these metastable, highly co-operative SSB-ssDNA complexes to their equilibrium, low co-operativity form is very slow at low concentrations of NaCl. At equilibrium, the SSB-ssDNA complexes seem to possess the same low degree of co-operativity (omega = 50 +/- 10) under all conditions tested. However, the highly co-operative mode of SSB binding, although metastable, may be important during non-equilibrium processes such as DNA replication. The possible relation between the two SSB binding modes, which differ in site size by a factor of two, and the high and low co-operativity complexes, which we report here, is discussed.  相似文献   

17.
The calf thymus DNA polymerase-alpha-primase complex purified by immunoaffinity chromatography catalyzes the synthesis of RNA initiators on phi X174 single-stranded viral DNA that are efficiently elongated by the DNA polymerase. Trace amounts of ATP and GTP are incorporated into products that are full length double-stranded circular DNAs. When synthetic polydeoxynucleotides are used as templates, initiation and DNA synthesis occurs with both poly(dT) and poly(dC), but neither initiation nor DNA synthesis was observed with poly(dA) and poly(dI) templates. Nitrocellulose filter binding and sucrose gradient centrifugation studies show that the DNA polymerase-primase complex binds to deoxypyrimidine polymers, but not to deoxypurine polymers. Using d(pA)-50 with 3'-oligo(dC) tails and d(pI)-50 with 3'-oligo(dT) tails, initiator synthesis and incorporation of deoxynucleotide can be demonstrated when the average pyrimidine sequence lengths are 8 and 4, respectively. These results suggest that purine polydeoxynucleotides are used as templates by the DNA polymerase only after initiation has occurred on the oligodeoxypyrimidine sequence and that the pyrimidine stretch required by the primase activity is relatively short. Analysis of initiator chain length with poly(dC) as template showed a series of oligo(G) initiators of 19-27 nucleotides in the absence of dGTP, and 5-13 nucleotides in the presence of dGTP. The chain length of initiators synthesized by the complex when poly(dT) or oligodeoxythymidylate-tailed poly(dI) was used can be as short as a dinucleotide. Analysis of the products of replication of oligo(dC)-tailed poly(dA) shows that initiator with chain length as low as 4 can be used for initiation by the polymerase-primase complex.  相似文献   

18.
Poly(dA).poly(dT) and DNA duplex with four or more adenine bases in a row exhibits a broad, solid-state structural premelting transition at about 35 degrees C. The low-temperature structure is correlated with the phenomena of "bent DNA." We have conducted temperature-dependent ultraviolet resonance Raman measurements of the structural transition using poly(dA).poly(dT) at physiological salt conditions, and are able to identify, between the high and low temperature limits, changes in the vibrational frequencies associated with the C4 carbonyl stretching mode in the thymine ring and the N6 scissors mode of the amine in the adenine ring of poly(dA).poly(dT). This work supports the model that the oligo-dA tracts' solid-state structural premelting transition is due to a set of cross-stand bifurcated hydrogen bonds between consecutive dA. dT pairs.  相似文献   

19.
Two-dimensional proton NMR studies are reported on the complementary d(C-A-T-G-T-G-T-A-C).d(G-T-A-C-epsilon A-C-A-T-G) nonanucleotide duplex (designated epsilon dA.dT 9-mer duplex) containing 1,N6-ethenodeoxyadenosine (epsilon dA), a carcinogen-DNA adduct, positioned opposite thymidine in the center of the helix. Our NMR studies have focused on the conformation of the epsilon dA.dT 9-mer duplex at neutral pH with emphasis on defining the alignment at the dT5.epsilon dA14 lesion site. The through-space NOE distance connectivities establish that both dT5 and epsilon dA14 adopt anti glycosidic torsion angles, are directed into the interior of the helix, and stack with flanking Watson-Crick dG4.dC15 and dG6.dC13 pairs. Furthermore, the d(G4-T5-G6).d(C13-epsilon A14-C15) trinucleotide segment centered about the dT5.epsilon dA14 lesion site adopts a right-handed helical conformation in solution. Energy minimization computations were undertaken starting from six different alignments of dT5(anti) and epsilon dA14(anti) at the lesion site and were guided by distance constraints defined by lower and upper bounds estimated from NOESY data sets on the epsilon dA.dT 9-mer duplex. Two families of energy-minimized structures were identified with the dT5 displaced toward either the flanking dG4.dC15 or the dG6.dC13 base pair. These structures can be differentiated on the basis of the observed NOEs from the imino proton of dT5 to the imino proton of dG4 but not dG6 and to the amino protons of dC15 but not dC13 that were not included in the constraints data set used in energy minimization. Our NMR data are consistent with a nonplanar alignment of epsilon dA14(anti) and dT5(anti) with dT5 displaced toward the flanking dG4.dC15 base pair within the d(G4-T5-G6).d(C13-epsilon A14-C15) segment of the epsilon dA.dT 9-mer duplex.  相似文献   

20.
We have examined the salt and temperature dependences of the equilibrium binding of the Escherichia coli single strand binding (SSB) tetramer to a series of oligodeoxythymidylates, dT(pT)N-1, with N = 16, 28, 35, 56 and 70. Absolute binding isotherms were obtained, based on the quenching of the intrinsic protein fluorescence upon formation of the complexes. The shorter oligonucleotides, with N = 16, 28 and 35, bind to multiple sites on the SSB tetramer and negative co-operativity is observed among these binding sites. We have quantitatively analyzed these isotherms, using a statistical thermodynamic ("square") model to obtain the intrinsic binding constant KN, and the negative co-operativity constant, sigma N. For all oligonucleotides, we find that KN decreases significantly with increasing concentration of monovalent salt, indicating a large electrostatic component to the free energy of the interaction (e.g. delta log KN/delta log [NaBr] = -2.7, -4.6 and -7.1 for N = 16, 35 and 70, respectively), with contributions from both cations and anions. For oligonucleotides that span two or more subunits, there is a significant unfavorable contribution to the binding free energy for each intersubunit crossing, with an accompanying uptake of anions. Therefore, the extent of anion uptake increases as the number of intersubunit crossings increase. There is a strong temperature dependence for the intrinsic binding of dT(pT)15, such that delta Ho = -26(+/- 3) kcal/mol dT(pT)15. Negative co-operativity exists under all solution conditions tested, i.e. sigma N less than 1, and this is independent of anion concentration and type. However, the negative co-operativity constant does decrease with decreasing concentration of cation. The dependence of sigma 16 on Na+ concentration indicates that an average of one sodium ion is taken up as a result of the negative co-operativity between two dT(pT)15 binding sites. These data and the lack of a temperature dependence for sigma 16 suggest that the molecular basis for the negative co-operativity is predominantly electrostatic and may be due to the repulsion of regions of single-stranded DNA that are required to bind in close proximity on an individual SSB tetramer.  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