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.
We examined the binding geometry of Co-meso-tetrakis (N-methyl pyridinium-4-yl)porphyrin, Co-meso-tetrakis (N-n-butyl pyridinium-4-yl)porphyrin and their metal-free ligands to poly[d(A-T)(2)] and poly[d(G-C)(2)] by optical spectroscopic methods including absorption, circular and linear dichroism spectroscopy, and fluorescence energy transfer technique. Signs of an induced CD spectrum in the Soret band depend only on the nature of the DNA sequence; all porphyrins exhibit negative CD when bound to poly[d(G-C)(2)] and positive when bound to poly[d(A-T)(2)]. Close analysis of the linear dichroism result reveals that all porphyrins exhibit outside binding when complexed with poly[d(A-T)(2)], regardless of the existence of a central metal and side chain. However, in the case of poly[d(G-C)(2)], we observed intercalative binding mode for two nonmetalloporphyrins and an outside binding mode for metalloporphyrins. The nature of the outside binding modes of the porphyrins, when complexed with poly[d(A-T)(2)] and poly[d(G-C)(2)], are quite different. We also demonstrate that an energy transfer from the excited nucleo-bases to porphyrins can occur for metalloporphyrins.  相似文献   

2.
The secondary structures of double-stranded poly[d(A-T)].poly[d(A-T)] in films have been studied by IR spectroscopy with three different counterions (Na+, Cs+, and Ni2+) and a wide variety of water content conditions (relative humidity between 100 and 47%). In addition to the A-, B-, C-, and D-form spectra, a new IR spectrum has been obtained in the presence of nickel ions. The IR spectra of Ni2+-poly[d(A-T)].poly[d(A-T)] films are analyzed by comparison with previously assigned IR spectra of left-handed poly[d(G-C)].poly[d(G-C)] and poly[d(A-C)].poly[d(G-T)], and it is possible to conclude that they reflect a Z-type structure for poly[d(A-T)].poly[d(A-T)]. The Z conformation has been favored by the high polynucleotide concentration, by the low water content of the films, and by specific interactions of the transition metal ions with the purine bases stabilized in a syn conformation. A structuration of the water hydration molecules around the double-stranded Ni2+-poly[d(A-T)].poly[d(A-T)] is shown by the presence of a strong sharp water band at 1615 cm-1.  相似文献   

3.
Poly[d(A-T).d(A-T)] and poly[d(G-C).d(G-C)], each dissolved in 0.1 M NaClO4, 5 mM cacodylic acid buffer, pH 6.8, experience inversion of their circular dichroism (CD) spectrum subsequent to the addition of Hg(ClO4)2. Let r identical to [Hg(ClO4)2]added/[DNA-P]. The spectrum of the right-handed form of poly[d(A-T).d(A-T)] turns into that of a seemingly left-handed structure at r greater than or equal to 0.05 while a similar transition is noted with poly[d(G-C).(G-C)] at r greater than or equal to 0.12. The spectral changes are highly cooperative in the long-wavelength region above 250 nm. At r = 1.0, the spectra of the two polymers are more or less mirror images of their CD at r = 0. While most CD bands experience red-shifts upon the addition of Hg(ClO4)2, there are some that are blue-shifted. The CD changes are totally reversible when Hg(II) is removed from the nucleic acids by the addition of a strong complexing agent such as NaCN. This demonstrates that mercury keeps all base pairs in register.  相似文献   

4.
Raman spectroscopy of Z-form poly[d(A-T)].poly[d(A-T)   总被引:3,自引:0,他引:3  
Helical structures of double-stranded poly[d(A-T)] in solution have been studied by Raman spectroscopy. While the classical right-handed conformation B-type spectra are obtained in the case of sodium chloride solutions, a Z-form Raman spectrum is observed by addition of nickel ions at high sodium concentration, conditions in which the inversion of the circular dichroic spectrum of poly[d(A-T)] is detected, similar to that observed for high-salt poly[d(G-C)] solutions [Bourtayre, P., Liquier, J., Pizzorni, L., & Taillandier, E. (1987) J. Biomol. Struct. Dyn. 5, 97-104]. The characterization of the Z-form spectrum of poly[d(A-T)] is proposed by comparison with previously obtained characteristic Raman lines of Z-form poly[d(G-C)] and poly[d(A-C)].poly[d(G-T)] solutions and of d(CG)3 and d(CGCATGCG) crystals [Thamann, T. J., Lord, R. C., Wang, A. H.-J., & Rich, A. (1981) Nucleic Acids Res. 9, 5443-5457; Benevides, J. M., Wang, A. H.-J., van der Marel, G. A., van Boom, J. H., Rich, A., & Thomas, G. J., Jr. (1984) Nucleic Acids Res. 14, 5913-5925]. Detailed spectroscopic data are presented reflecting the reorientation of the purine-deoxyribose entities (C2'-endo/anti----C3'-endo/syn), the modification of the phosphodiester chain, and the adenosine lines in the 1300-cm-1 region. The role played by the hydrated nickel ions in the B----Z transition is discussed.  相似文献   

5.
The binding of propidium to poly(dA).poly(dT) [poly(dA.dT)] and to poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2]] has been compared under a variety of solution conditions by viscometric titrations, binding studies, and kinetic experiments. The binding of propidium to poly[d(A-T)2] is quite similar to its binding to calf thymus deoxyribonucleic acid (DNA). The interaction with poly(dA.dT), however, is quite unusual. The viscosity of a poly(dA.dT) solution first decreases and then increases in a titration with propidium at 18 degrees C. The viscosity of poly[d(A-T)2] shows no decrease in a similar titration. Scatchard plots for the interaction of propidium with poly(dA.dT) show the classical upward curvature for positive cooperativity. The curvature decreases as the temperature is increased in binding experiments. A van't Hoff plot of the observed binding constants yields an apparent positive enthalpy of approximately +6 kcal/mol for the propidium-poly(dA.dT) interaction. Propidium binding to poly[d(A-T)2] shows no evidence for positive cooperativity, and the enthalpy change for the reaction is approximately -9 kcal/mol. Both the magnitude of the dissociation constants and the effects of ionic strength are quite similar for the dissociation of propidium from poly(dA-T)2] and from poly[d(A-T)2], suggesting that the intercalated states are similar for the two complexes. The observed association reactions, under pseudo-first-order conditions, are quite different. Plots of the observed pseudo-first-order association rate constant vs. polymer concentration have much larger slopes for propidium binding to poly[d(A-T)2] than to poly(dA.dT).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The binding of lac repressor to poly d(A-T) and poly d(G-C) has been studied using circular dichroism. The results indicate that the binding induces the same conformational change of both polynucleotides and perturbs the same number of nucleic acid bases (28 bases). It is shown that in 0.1 M phosphate buffer the CD measurement can be used to determine the binding constant of lac repressor to poly d(A-T). Competition experiments performed at various salt concentrations show that the stronger interaction of lac repressor for poly d(A-T) than for poly d(G-C) is based on difference in the dissociation rate of the complexes whereas the association rate for both polymers are similar.  相似文献   

7.
The binding of polyamines, including spermidine ( 1 ) and spermine ( 2 ), to poly[d(G-C) · d(G-C) ] was probed using spectroscopic studies of anthracene-9-carbonyl-N1-spermine ( 3 ); data from normal absorption, linear dichroism (LD), and circular dichroism (CD) are reported. Ligand LD and CD for transitions located in the DNA region of the spectrum were used. The data show that 3 binds to DNA in a manner characteristic of both its amine and polycyclic aromatic parts. With poly [(dG-dC) · (dG-dC)], binding modes are occupied sequentially and different modes correspond to different structural perturbations of the DNA. The most stable binding mode for 3 with poly[d(G-C) · d(G-C)] has a site size of 6 ± 1 bases, and an equilibrium binding constant of (2.2 ± 1.1) × 107 M?1 with the anthracene moiety intercalated. It dominates the spectra from mixing ratios of approximately 133:1 until 6:1 DNA phosphate: 3 is reached. The analogous data for poly [d(A-T) · d(A-T)] between mixing ratios 36:1 and 7:1 indicates a site size of 8.3 ± 1.1 bases and an equilibrium binding constant of (6.6 ± 3.3) × 105 M?1. Thus, 3 binds preferentially to poly [d(G-C) · d(G-C)] at these concentrations. © 1994 John Wiley & Sons, Inc.  相似文献   

8.
Equilibrium binding experiments using fluorescence and absorption techniques have been performed throughout a wide concentration range (1 nM to 30 microM) of the dye Hoechst 33258 and several DNAs. The most stable complexes found with calf thymus DNA, poly[d(A-T)], d(CCGGAATTCCGG), and d(CGCGAATTCGCG) all have dissociation constants in the range (1-3) X 10(-9) M-1. Such complexes on calf thymus DNA occur with a frequency of about 1 binding site per 100 base pairs, and evidence is presented indicating a spectrum of sequence-dependent affinities with dissociation constants extending into the micromolar range. In addition to these sequence-specific binding sites on the DNA, the continuous-variation method of Job reveals distinct stoichiometries of dye-poly[d(A-T)] complexes corresponding to 1, 2, 3, 4, and 6 dyes per 5 A-T base pairs and even up to 1 and 2 (and possibly more) dyes per backbone phosphate. Models are suggested to account for these stoichiometries. With poly[d(G-C)] the stoichiometries are 1-2 dyes per 5 G-C pairs in addition to 1 and 2 dyes per backbone phosphate. Thermodynamic parameters for formation of the tightest binding complex between Hoechst 33258 and poly[d(A-T)] or d-(CCGGAATTCCGG) are determined. Hoechst 33258 binding to calf thymus DNA, chicken erythrocyte DNA, and poly[d(A-T)] exhibits an ionic strength dependence similar to that expected for a singly-charged positive ion. This ionic strength dependence remains unchanged in the presence of 25% ethanol, which decreases the affinity by 2 orders of magnitude. In addition, due to its strong binding, Hoechst 33258 easily displaces several intercalators from DNA.  相似文献   

9.
The interaction of Ethidium Bromide (EtBr) with double-stranded (ds-) and single-stranded (ss-) poly[d(A-T)] was studied in different ionic strengths solutions. Optical spectroscopy and Scatchard analysis results indicate that the ligand interacts to both helix and coiled structures of the polynucleotide by "strong" and "weak" binding modes. The association parameters (binding constant -K- and the number of nucleotides corresponding to a binding site -n) of the strong type of interaction were found to be independent of Na+ concentration. Weak interaction occurs at low ionic strength and/or high EtBr concentration. Estimated binding parameters of EtBr with ss- and ds-polynucleotide are in good agreement with those for EtBr-B-DNA complexes. Data obtained provided an evidence for a stacking interaction of EtBr with single stranded poly[d(A-T)].  相似文献   

10.
31P- and 1H-nmr and laser Raman spectra have been obtained for poly[d(G-T)]·[d(C-A)] and poly[d(A-T)] as a function of both temperature and salt. The 31P spectrum of poly[d(G-T)]·[d(C-A)] appears as a quadruplet whose resonances undergo separation upon addition of CsCl to 5.5M. 1H-nmr measurements are assigned and reported as a function of temperature and CsCl concentration. One dimensional nuclear Overhauser effect (NOE) difference spectra are also reported for poly[d(G-T)]·[d(C-A)] at low salt. NOE enhancements between the H8 protons of the purines and the C5 protons of the pyrimidines, (H and CH3) and between the base and H-2′,2″ protons indicate a right-handed B-DNA conformation for this polymer. The NOE patterns for the TH3 and GH1 protons in H2O indicate a Watson–Crick hydrogen-bonding scheme. At high CsCl concentrations there are upfield shifts for selected sugar protons and the AH2 proton. In addition, laser Raman spectra for poly[d(A-T)] and poly[d(G-T)]·[d(C-A)] indicate B-type conformations in low and high CsCl, with predominantly C2′-endo sugar conformations for both polymers. Also, changes in base-ring vibrations indicate that Cs+ binds to O2 of thymine and possibly N3 of adenine in poly[d(G-T)]·[d(C-A)] but not in poly[d(A-T)]. Further, 1H measurements are reported for poly[d(A-T)] as a function of temperature in high CsCl concentrations. On going to high CsCl there are selective upfield shifts, with the most dramatic being observed for TH1′. At high temperature some of the protons undergo severe changes in linewidths. Those protons that undergo the largest upfield shifts also undergo the most dramatic changes in linewidths. In particular TH1′, TCH3, AH1′, AH2, and TH6 all undergo large changes in linewidths, whereas AH8 and all the H-2′,2″ protons remain essentially constant. The maximum linewidth occurs at the same temperature for all protons (65°C). This transition does not occur for d(G-T)·d(C-A) at 65°C or at any other temperature studied. These changes are cooperative in nature and can be rationalized as a temperature-induced equilibrium between bound and unbound Cs+, with duplex and single-stranded DNA. NOE measurements for poly[d(A-T)] indicate that at high Cs+ the polymer is in a right-handed B-conformation. Assignments and NOE effects for the low-salt 1H spectra of poly[d(A-T)] agree with those of Assa-Munt and Kearns [(1984) Biochemistry 23 , 791–796] and provide a basis for analysis of the high Cs+ spectra. These results indicate that both polymers adopt a B-type conformation in both low and high salt. However, a significant variation is the ability of the phosphate backbone to adopt a repeat dependent upon the base sequence. This feature is common to poly[d(G-T)]·[d(C-A)], poly[d(A-T)], and some other pyr–pur polymers [J. S. Cohen, J. B. Wouten & C. L Chatterjee (1981) Biochemistry 20 , 3049–3055] but not poly[d(G-C)].  相似文献   

11.
Binding configurations and equilibria of intercalation complexes formed by the novel anthracycline drug, 2-fluoro-4-demethoxydaunomycin (2FD), with the decanucleotides d(G-C)5 and d(A-T)5 have been studied by 19F-NMR spectroscopy. The 19F chemical shift of 2FD bound to d(A-T)5 was approximately 1.5 ppm downfield of that observed for 2FD bound to d(G-C)5. By mixing equimolar amounts of aqueous d(G-C)5, d(A-T)5 and 2FD, the distribution of drug between the nucleotides was followed by observing relative peak intensities and showed no G-C or A-T binding preference at room temperature. It was shown that each decanucleotide duplex bound three 2FD molecules, giving a neighbour exclusion parameter, n, of n = 3 for this drug. The stoichiometric complexes, which we denote by [d(A-T)5][2FD]3 and [d(G-C)5][2FD]3, were also purified and isolated in this study.  相似文献   

12.
We report the temperature and salt dependence of the volume change (DeltaVb) associated with the binding of ethidium bromide and netropsin with poly(dA).poly(dT) and poly[d(A-T)].poly[d(A-T)]. The DeltaV(b) of binding of ethidium with poly(dA).poly(dT) was much more negative at temperatures approximately 70 degrees C than at 25 degrees C, whereas the difference is much smaller in the case of binding with poly[d(A-T)].poly[d(A-T)]. We also determined the volume change of DNA-drug interaction by comparing the volume change of melting of DNA duplex and DNA-drug complex. The DNA-drug complexes display helix-coil transition temperatures (Tm several degrees above those of the unbound polymers, e.g., the Tm of the netropsin complex with poly(dA)poly(dT) is 106 degrees C. The results for the binding of ethidium with poly[d(A-T)].poly[d(A-T)] were accurately described by scaled particle theory. However, this analysis did not yield results consistent with our data for ethidium binding with poly(dA).poly(dT). We hypothesize that heat-induced changes in conformation and hydration of this polymer are responsible for this behavior. The volumetric properties of poly(dA).poly(dT) become similar to those of poly[d(A-T)].poly[d(A-T)] at higher temperatures.  相似文献   

13.
gp32 I is a protein with a molecular weight of 27 000. It is obtained by limited hydrolysis of T4 gene 32 coded protein, which is one of the DNA melting proteins. gp32 I itself appears to be also a melting protein. It denatures poly[d(A-T)].poly[d(A-T)] and T4 DNA at temperatures far (50-60 degrees C) below their regular melting temperatures. Under similar conditions gp32 I will denature poly[d(A-T).poly[d(A-T)] at temperatures approximately 12 degrees C lower than those measured for the intact gp32 denaturation. For T4 DNA gp32 shows no melting behavior while gp32 I shows considerable denaturation (i.e., hyperchromicity) even at 1 degree C. In this paper the denaturation of poly[d(A-T)].poly[d(A-T)] and T4 DNA by gp32 I is studied by means of circular dichroism. It appears that gp32 I forms a complex with poly[d(A-T)]. The conformation of the polynucleotide in the complex is equal to that of one strand of the double-stranded polymer in 6 M LiCl. In the gp32 I DNA complex formed upon denaturation of T4 DNA, the single-stranded DNA molecule has the same conformation as one strand of the double-strand T4 DNA molecule in the C-DNA conformation.  相似文献   

14.
H H Klump  E Schmid    M Wosgien 《Nucleic acids research》1993,21(10):2343-2348
The conformational change for the alternating purine-pyrimidine polydeoxyribonucleotides i.e. poly d(A-T), poly d(G-C), and poly d(A-C) poly d(G-T) from a right-handed conformation at room temperature to the left-handed Z-DNA like double helix at elevated temperatures has been studied by UV spectroscopy, Raman spectroscopy, and by adiabatic differential scanning microcalorimetry (DSC) in the presence of Na+ and Mg2+ or Ni2+ respectively as counterions. The differential UV spectra reveal through a hyperchromic shift at around 280nm and a hypochromic shift at 260nm that a conformational change to the left-handed conformation occurs. The Raman spectra clearly show characteristic changes, a drastic decrease of the band at 680cm-1 and the appearance of a new band at 628cm-1, due to the change of the purine bases to the syn conformation upon inversion of the helix-handedness. The course of the transition as function of temperature can be followed quantitatively by plotting the change in the excess heat capacity vs. temperature. The transition enthalpy delta H for the B- to Z-DNA transition per mole base pairs (mbp) amounts to 2.0 +/- 0.2kcal for poly d(G-C), to 4.0 +/- 0.4kcal for poly d(A-T), and to 3.1 +/- 0.3kcal for poly d(A-C) poly d(G-T). The enthalpy change due to the Z-DNA to coil transitions (per mole base pairs) amounts to 11kcal for poly d(G-C), 10.5kcal for poly d(A-T) and 11.3kcal for poly d(A-C) poly d(G-T).  相似文献   

15.
F A Tanious  S F Yen  W D Wilson 《Biochemistry》1991,30(7):1813-1819
The interaction of a symmetric naphthalene diimide with alkylamino substituents at each imide position was investigated with the alternating sequence polymers, poly[d(A-T)]2 and poly[d(G-C)]2. Spectrophotometric binding studies indicate strong binding of the diimide to both sequences although the GC binding constant is 20-25 times larger than the AT binding constant. Analysis of the effects of salt concentration on the binding equilibria shows that the diimide forms two ion pairs in its complex with both polymers as expected for a simple dication. Stopped-flow kinetics experiments demonstrate that the diimide both associates and dissociates from DNA more slowly than classical intercalators with similar binding constants. Analysis of salt concentration effects on dissociation kinetics rate constants (kd) reveals that slopes in log kd versus log [Na+] plots are only approximately half the value obtained for classical dicationic intercalators that have both charged groups in the same groove. These kinetics results support a threading intercalation model, with one charged diimide substituent in each of the DNA grooves rather than with both side chains in the same groove, for the diimide complex with DNA. In the rate-determining step of the mechanism for dissociation of a threading complex only one ion pair is broken; the free side chain can then slide between base pairs to put both diimide side chains in the same groove, and this is followed by rapid full dissociation of the diimide. This sequential release of ion pairs makes the dissociation slope for dicationic threading intercalators more similar to the slope for classical monocationic intercalating ligands.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Near UV CD spectra, UV absorption spectra and their first derivatives have been recorded on poly d(A-T).poly d(A-T) solutions in presence of high NaCl concentration and various amounts of NiCl2. Comparison of the results presented here with those obtained for poly d(G-C).poly d(G-C) and poly d(A-C).poly d(G-T) in comparable conditions, and the I.R. and Raman data on poly d(A-T).poly d(A-T), allows us to assign the new spectra to the Z conformation of poly d(A-T).poly d(A-T) in solution. The mechanism by which nickel ions induce the B----Z interconversion in the presence of high NaCl concentration is discussed.  相似文献   

17.
Interaction of a macrocyclic bisacridine with DNA   总被引:1,自引:0,他引:1  
The binding of the macrocycle SDM to DNA was investigated by visible spectroscopy, stopped-flow kinetics, and NMR spectroscopy. SDM is composed of two 9-aminoacridines linked via the amino groups by a spermine side chain and via the 4-positions by a N,N'-[(methylthio)ethyl]succinamide side chain [Zimmerman, S. C., Lamberson, C. R., Cory, M., & Fairley, T. A. (1989) J. Am. Chem. Soc. 111, 6805-6809]. The visible spectrum of SDM bound to poly[d(A-T)]2 or poly[d(G-C)]2 is red-shifted relative to the spectrum of SDM alone and displays considerable hypochromicity. Results from titrations of SDM with polymer indicate a binding site size of three base pairs per macrocycle. The dissociation constant for SDM bound to either poly[d(A-T)]2 or poly[d(G-C)]2 is an order of magnitude lower than that for a similar bisacridine linked only by a spermine side chain. In addition, the dependence of the dissociation constant on ionic strength is significantly reduced. NMR studies of SDM complexes with poly[d(A-T)]2 or a tetramer, d(CGCG)2, show that intercalation is the mode of binding. The magnitudes of the chemical shift differences for SDM aromatic protons in the free and bound states support intercalation with the acridine ring systems essentially parallel to the long axis of the base pairs. Cross peaks from NOESY spectra of the SDM complex with d(CGCG)2 further support this mode of binding and provide information on the structure of the complex. The results are analyzed for consistency with each of three binding models: (i) bisintercalation with the two side chains in the same groove; (ii) bisintercalation according to the neighbor-exclusion principle with the two side chains in opposite grooves; and (iii) bisintercalation with two side chains in opposite grooves but with violation of the neighbor-exclusion principle. Model i is found to be unlikely on the basis of all evidence obtained, including preliminary modeling studies. Both models ii and iii can be reconciled with the experimental evidence and from a modeling standpoint are energetically feasible.  相似文献   

18.
A F Corin  T M Jovin 《Biochemistry》1986,25(14):3995-4007
The delayed fluorescence properties of proflavin have been exploited in studies of the excited-state binding kinetics of the dye to poly[d(A-T)] and its brominated analogue poly[d(A-br5U)] at room temperature and pH 7. The two analyzed luminescence decay times of the DNA-dye complex are dependent on the total nucleic acid concentration. This dependence is shown to reflect a temporal coupling of the intrinsic delayed emission decay rates with the dynamic chemical kinetic binding processes in the excited state. Temperature-jump kinetic studies conducted on the brominated polymer and corresponding information on poly[d(A-T)] from a previous study [Ramstein, J., Ehrenberg, M., & Rigler, R. (1980) Biochemistry 19, 3938-3948] provide complementary information about the ground state. In the ground state, the poly[d(A-T)]-proflavin complex has one chemical relaxation time, which reaches a plateau at high DNA concentrations. The brominated DNA-dye complex exhibits two relaxation times: a faster relaxation mode that behaves similarly to that for the unhalogenated DNA and a slower relaxation mode that is apparent at high DNA concentrations. The ground-state kinetic data are analyzed in terms of two alternative models incorporating series and parallel reaction schemes. The former consists of two sequential binding steps--a fast bimolecular process followed by a monomolecular step--while the latter consists of two coupled bimolecular steps. A similar analysis for the excited-state data yields reasonable kinetic constants only for the series model, which, in accordance with previous proposals for acridine intercalators, consists of a fast outside binding step followed by intercalation of the dye. A comparison of the ground- and excited-state kinetic parameters reveals that the external binding process is much stronger and the intercalation is much weaker in the excited state. That the excited-state data are only consistent with the series model suggests that delayed luminescence studies may provide a general tool for distinguishing between the two kinetic mechanisms. In particular, we demonstrate the use of delayed luminescence spectroscopy as a tool for probing dynamic DNA-ligand interactions in solution.  相似文献   

19.
We have undertaken a search for mammalian DNA-binding proteins that enhance the activity of DNA polymerases in a template sequence-specific fashion. In this paper, we report the extensive purification and characterization of a new DNA-binding protein from rabbit liver that selectively stimulates DNA polymerases to copy synthetic poly[d(G-C)] and the poly(dC) strand of poly(dC).poly(dG) as well as single-stranded natural DNA that contains stretches of oligo(dC). The enhancing protein, a polypeptide of 65 kDa designated factor C, stimulates the copying of the two synthetic templates by Escherichia coli DNA polymerase I, Micrococcus luteus polymerase, and eukaryotic DNA polymerases alpha and beta, but not by avian myeloblastosis virus polymerase. Factor C, however, does not affect utilization by these polymerases of the poly(dG) strand of poly(dC).poly(dG), of poly(dC) primed by oligo(dG), or of poly(dA).poly(dT) and poly[d(A-T)]. With polymerase I, Michaelis constants (Km) of poly[d(G-C)] and of the poly(dC) strand of poly(dC).poly(dG) are decreased by factor C 37- and 4.7-fold, respectively, whereas maximum velocity (Vmax) remains unchanged. By contrast, neither the Km value of the poly(dG) strand of poly(dC).poly(dG) nor the Vmax value with this template is altered by factor C. Rates of copying of activated DNA, denatured DNA, or singly primed M13 DNA are not affected significantly by factor C. However, primer extension analysis of the copying of recombinant M13N4 DNA that contains runs of oligo(dC) within an inserted thymidine kinase gene shows that factor C increases processivity by specifically augmenting the efficiency at which polymerase I traverses the oligo(dC) stretches. Direct binding of factor C to denatured DNA is indicated by retention of the protein-DNA complex on columns of DEAE-cellulose. Binding of factor C to poly[d(G-C)] is demonstrated by the specific adsorption of the enhancing protein to columns of poly[d(G-C)]-Sepharose. We propose that by binding to poly[d(G-C)] and to poly(dC).poly(dG), factor C enables tighter binding of some DNA polymerases to these templates and facilitates enzymatic activity.  相似文献   

20.
B C Sang  D M Gray 《Biochemistry》1987,26(23):7210-7214
Circular dichroism (CD) data indicated that fd gene 5 protein (G5P) formed complexes with double-stranded poly(dA.dT) and poly[d(A-T).d(A-T)]. CD spectra of both polymers at wavelengths above 255 nm were altered upon protein binding. These spectral changes differed from those caused by strand separation. In addition, the tyrosyl 228-nm CD band of G5P decreased more than 65% upon binding of the protein to these double-stranded polymers. This reduction was significantly greater than that observed for binding to single-stranded poly(dA), poly(dT), and poly[d(A-T)] but was similar to that observed for binding of the protein to double-stranded RNA [Gray, C.W., Page, G.A., & Gray, D.M. (1984) J. Mol. Biol. 175, 553-559]. The decrease in melting temperature caused by the protein was twice as great for poly[d(A-T).d(A-T)] as for poly(dA.dT) in 5 mM tris(hydroxymethyl)aminomethane hydrochloride (Tris-HCl), pH 7. Upon heat denaturation of the poly(dA.dT)-G5P complex, CD spectra showed that single-stranded poly(dA) and poly(dT) formed complexes with the protein. The binding of gene 5 protein lowered the melting temperature of poly(dA.dT) by 10 degrees C in 5 mM Tris-HCl, pH 7, but after reducing the binding to the double-stranded form of the polymer by the addition of 0.1 M Na+, the melting temperature was lowered by approximately 30 degrees C. Since increasing the salt concentration decreases the affinity of G5P for the poly(dA) and poly(dT) single strands and increases the stability of the double-stranded polymer, the ability of the gene 5 protein to destabilize poly(dA.dT) appeared to be significantly affected by its binding to the double-stranded form of the polymer.  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