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 15 毫秒
1.
A sensitive ESR method which allows a direct quantitative determination of nucleic acid binding affinities of proteins under physiologically relevant conditions has been applied to the gene 5 protein of bacteriophage fd. This was achieved with two spin-labeled nucleic acids, (ldT, dT)n and (lA,A)n, which served as macro-molecular spin probes in ESR competition experiments. With the two different macromolecular spin probes, it was possible to determine the relative apparent affinity constants, Kapp, over a large affinity domain. In 20 mM Tris X HCl (pH 8.1), 1 mM sodium EDTA, 0.1 mM dithiothreitol, 10% (w/v) glycerol, 0.05% Triton, and 125 mM NaCl, the following affinity relationship was observed: K(dT)napp = 10(3) KfdDNAapp = 2 X 10(4) K(A)napp = 6.6 X 10(4) KrRNAapp = 1.5 X 10(5) KR17RNAapp. Increasing the [NaCl] from 125 to 200 mM caused considerably less tight binding of gene 5 protein to (lA,A)n, and a typical cooperative binding isotherm was observed, whereas at the lower [NaCl] used for the competition experiments, the binding was essentially stoichiometric. A computer fit of the experimental titration data at 200 mM NaCl gave an intrinsic binding constant, Kint, of 1300 M-1 and a cooperativity factor, omega, of 60 (Kint omega = Kapp) for (lA,A)n.  相似文献   

2.
The binding properties of the Escherichia coli encoded single strand binding protein (SSB) to a variety of synthetic homopolynucleotides, as well as to single stranded M13 DNA, have been examined as a function of the NaCl concentration (25.0 degrees C, pH 8.1). Quenching of the intrinsic tryptophan fluorescence of the SSB protein by the nucleic acid is used to monitor binding. We find that the site size (n) for binding of SSB to all single stranded nucleic acids is quite dependent on the NaCl concentration. For SSB-poly(dT), n = 33 +/- 3 nucleotides/tetramer below 10 mM NaCl and 65 +/- 5 nucleotides/tetramer above 0.20 M NaCl (up to 5 M). Between 10 mM and 0.2 M NaCl, the apparent site size increases continuously with [NaCl]. The extent of quenching of the bound SSB fluorescence by poly(dT) also displays two-state behavior, 51 +/- 3% quenching below 10 mM NaCl and 83 +/- 3% quenching at high [NaCl] (greater than 01.-0.2 M NaCl), which correlates with the observed changes in the occluded site size. On the basis of these observations as well as the data of Krauss et al. (Krauss, G., Sindermann, H., Schomburg, U., and Maass, G. (1981) Biochemistry 20, 5346-5352) and Chrysogelos and Griffith (Chrysogelos, S., and Griffith, J. (1982) Proc. Natl. Acad. Sci. U. S. A. 79,5803-5807) we propose a model in which E. coli SSB binds to single stranded nucleic acids in two binding modes, a low salt mode (n = 33 +/- 3), referred to as (SSB)33, in which the nucleic acid interacts with only two protomers of the tetramer, and one at higher [NaCl], n = 65 +/- 5, (SSB)65, in which the nucleic acid interacts with all 4 protomers of the tetramer. At intermediate NaCl concentrations a mixture of these two binding modes exists which explains the variable site sizes and other apparent discrepancies previously reported for SSB binding. The transition between the two binding modes is reversible, although the kinetics are slow, and it is modulated by NaCl concentrations within the physiological range. We suggest that SSB may utilize both binding modes in its range of functions (replication, recombination, repair) and that in vivo changes in the ionic media may play a role in regulating some of these processes.  相似文献   

3.
Sequence-specific spin-labeled oligodeoxynucleotides with conformation-sensitive electron paramagnetic resonance (EPR) signals are synthesized and examined as solution-phase nucleic acid hybridization probes. Either a proxyl or tempo ring linked to the C(5) position of deoxyuridine (dU) by a nonrigid two-atom methylamino tether is incorporated within 15-mers by phosphotriester chemistry yielding stable spin-labeled probes with distinctive EPR specific activity (AEPR) values. The AEPR is greater for a proxyl-labeled than for a tempo-labeled probe and is consistent with EPR data of enzymatically labeled 26-mers [Bobst, A. M., Pauly, G. T., Keyes, R. S., and Bobst, E. V. (1988) FEBS Lett. 228, 33-36], after normalizing for percent labeling. The spectral characteristics of the free probes and the probe/target complexes are similar to those of enzymatically spin-labeled nucleic acids containing a different nonrigid two-atom-tethered spin label [Bobst, A. M., Kao, S.-C., Toppin, R. C., Ireland, J. C., and Thomas, I. E. (1984) J. Mol. Biol. 173, 63-70]. The presence of target DNA is detected in solution by EPR spectroscopy and the assay is based on the characteristic line-shape change associated with hybridization. The EPR spectra of free and bound probe reflect little interference from changes in global dynamics of the probe, and the line-shape change upon complexation results primarily from a change in local base dynamics. The presence or absence of hybridization can be detected in a loop-gap resonator with about 1 pmol of spin-labeled 15-mer within minutes.  相似文献   

4.
Electron paramagnetic resonance (EPR) spectra of the two-atom-tethered six-membered ring thymidylate spin label (DUMTA) incorporated into duplexes of different sizes were found to display a helix length dependence and a local-order parameter S = 0.32 +/- 0.01 for B-DNA based on the dynamic cylinder model (Keyes, R. S., and A. M. Bobst. 1995. Detection of internal and overall dynamics of a two-atom-tethered spin-labeled DNA. Biochemistry. 34:9265-9276). This sensitivity to size, which reflects global tumbling, is now reported for the more flexible five-atom-tethered five-membered ring thymidylate spin label (DUAP) that can be readily incorporated enzymatically and sequence specifically into nucleic acids of different sizes. The DUAPs containing B-DNA systems were simulated with the same dynamic cylinder model, giving S = 0.20 +/- 0.01 for the more flexibly tethered spin label. This shows that S is dependent on tether length but not on global motion. An analysis with the same motional model of the B-Z transition in a (dG-dC)n polymer containing the five-atom-tethered six-membered ring cytidylate spin label (DCAT) (Strobel, O. K., R. S. Keyes, and A. M. Bobst. 1990b. Base dynamics of local Z-DNA conformations as detected by electron paramagnetic resonance with spin-labeled deoxycytidine analogues. Biochemistry. 29:8522-8528) revealed an increase in S from 0.15 +/- 0.01 to 0.26 +/- 0.01 in response to the B- to Z-DNA transition. This indicates that S is not only sensitive to tether length, but also to conformational changes in DNA. Both the DUAP- and the DCAT-labeled systems were also simulated with a base disk model. From the DUAP spectral series, the perpendicular component of the correlation time tau perpendicular describing the spin-labeled base diffusion was found to be sensitive to global tumbling, confirming earlier results obtained with DUMTA. The DCAT polymer results demonstrated that tau perpendicular monitors a conformational change from B- to Z-DNA, indicating that tau perpendicular is also sensitive to local base dynamics. These results confirm that the dynamics of five-atom-tethered nitroxides are coupled to the nucleic acid dynamics and, as with two-atom-tethered spin labels, can be characterized by S and tau perpendicular. The analyses of both spin-labeled systems provide good evidence for spin-labeled base motions within double-stranded DNA occurring on the nanosecond time scale, and establish that both labels can be used to monitor changes in global tumbling and local order parameter due to variations in DNA conformation and protein-DNA interactions.  相似文献   

5.
O K Strobel  R S Keyes  A M Bobst 《Biochemistry》1990,29(37):8522-8528
Conformation detection and base dynamics of spin-labeled Z-DNA have been investigated by electron paramagnetic resonance (EPR) spectroscopy. The two synthesized and characterized probes used in this study were C(5)-nitroxide-labeled 2'-deoxycytidine 5'-triphosphates, pppDCAT and pppDCAVAT, which serve as suitable substrates for Micrococcus luteus DNA polymerase. Enzymatic incorporation of these probes into (dG-dC)n yields the EPR-active alternating copolymers (dG-dC,DCAT)n and (dG-dC,-DCAVAT)n. These polymers assume typical B- and Z-DNA conformations under respective low (0.1 M NaCl) and high (4.5 M NaCl) salt conditions, as evidenced by their UV-circular dichroism spectra. The EPR line shape of (dG-dC,DCAT)n in Z-form is unique and significantly different from the B-form EPR spectrum. A similar observation is made for (dG-dC,DCAVAT)n. Thus, the EPR line shapes of these spin-labeled DNAs are indicative of their local conformations. The EPR spectra, analyzed with a previously published motional model [Kao, S.-C., Polnaszek, C.F., Toppin, C.R., & Bobst, A.M. (1983) Biochemistry 22, 5563-5568], indicate tau perpendicular values of 4 and 7 ns for the B- and Z-forms, respectively. Therefore, the base dynamics of Z-DNA are about two times slower than in B-DNA.  相似文献   

6.
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   

7.
In an extension of earlier studies on the Escherichia coli plasmid-encoded single-stranded DNA-binding proteins pIP71a SSB, F SSB and R64 SSB [Khamis, M. I., Casas-Finet, J. R., Maki, A. H., Ruvolo, P. P. & Chase, J. W. (1987) Biochemistry 26, 3347-3354; Casas-Finet, J. R., Khamis, M. I., Maki, A. H., Ruvolo, P. P. & Chase, J. W. (1987) J. Biol. Chem. 262, 8574-8593], we have investigated the binding of pIP231a SSB to natural and heavy-atom-derivatized single-stranded homopolynucleotides. Fluorimetric equilibrium binding isotherms indicate that pIP231a SSB has a greater solubility at low ionic strength than any other plasmid SSB protein investigated. Furthermore, its complex with mercurated poly(uridylic acid) [poly(Hg5U)] shows a greater resistance to disruption by salt than the other plasmid SSB complexes. Essentially complete binding of pIP231a SSB to poly(Hg5U) could be achieved, and time-resolved optically detected triplet-state magnetic resonance (ODMR) techniques could be applied to the complex. These methods allowed complete resolution of the three Trp chromophores of pIP231a SSB. Comparison of wavelength-selected ODMR results with those obtained for the poly(Hg5U) complex of a point-mutated chromosomal ssb gene product (Eco SSB) carrying substitutions of Phe for Trp [Khamis, M. I., Casas-Finet, J. R., Maki, A. H., Murphy, J. B. & Chase, J. W. (1987) J. Biol. Chem. 262, 10938-10945] confirm that Trp40 and Trp54 of pIP231a SSB are stacked in the complex, while Trp88 is not. This is the same distribution of stacked Trp residues found in Eco SSB. These results are confirmed further by specific effects observed on the ODMR signals of pIP231a SSB upon binding to poly(Br5U) and poly(dT), which are known to be caused by the stacking of Trp54 with nucleic acid bases.  相似文献   

8.
Gene 5 protein-DNA complex: modeling binding interactions   总被引:1,自引:0,他引:1  
A helical (not toroidal) complex consisting of eight gene 5 protein dimers per turn is proposed for the extension of DNA from dimer to dimer using known bond length constraints, postulated protein-nucleic acid interactions (determined from NMR and chemical modification studies), other physical properties of the complex, and data from electron micrographs. The binding channel has been dictated by these known parameters and the relative ease of geometrically fitting these constituents. This channel is different from that previously reported by other modelers. The channel lies underneath the long arm "claw-like" extension of the monomer, so that it rests inside the outer surface of the protein complex. An explanation is proposed for the two binding modes, n = 4 (the predominate mode) and n = 3, based on the weak binding interaction of Tyrosine 34. Also, the site of the less mobile nucleic acid base as reported from ESR studies (S.-C. Kao, E.V. Bobst, G.T. Pauly and A.M. Bobst, J. Biom. Struc. Dyn. 3,261 (1985)) is postulated as involving the fourth nucleotide, and this particular base is stacked between Tyrosine 34 and Phenylalanine 73'.  相似文献   

9.
T M Lohman 《Biochemistry》1984,23(20):4665-4675
The dissociation kinetics of bacteriophage T4 coded gene 32 protein-single-stranded nucleic acid complexes have been examined as a function of monovalent salt concentration, temperature, and pH in order to investigate the details of the dissociation of cooperatively bound protein. Fluorescence stopped-flow techniques were used, and irreversible dissociation was induced by a combination of [NaCl] jumps and mixing with excess nucleic acid competitor. This made it possible to directly investigate the irreversible dissociation process over a wide range of NaCl concentrations [e.g., from 50 mM to 0.60 M for the gene 32 protein-poly(A) complex], in the absence of reassociation. Over the entire salt range, the only dissociable species observed is the singly contiguously bound gene 32 protein which dissociates from the ends of protein clusters. However, the [NaCl] dependence of the dissociation rate constant suggests that two competing pathways exist for dissociation of cooperatively bound gene 32 protein from the ends of protein clusters. At high monovalent salt concentrations, dissociation is dominated by a single-step process, with log ke/log [NaCl] = 6.5 +/- 0.5; i.e., the dissociation rate constant increases with increasing NaCl concentration due to the uptake of approximately six monovalent ions upon dissociation. This indicates that singly contiguous protein dissociates directly into solution. However, at much lower [NaCl] the data suggest that gene 32 protein, when bound at the end of a protein cluster, dissociates by first sliding off the end to form a noncooperatively bound intermediate which subsequently dissociates. A quantitative model which incorporates the sliding pathway [Berg, O. G., Winter, R. B., & von Hippel, P. H. (1981) Biochemistry 20, 6929-6948] in the dissociation mechanism fits the data reasonably well and suggests that noncooperatively bound monomers of gene 32 protein may be capable of one-dimensional translocation along single-stranded nucleic acids as suggested by independent kinetic data on the association reaction [Lohman, T. M., & Kowalczykowski, S. C. (1981) J. Mol. Biol. 152, 67-109]. It is also observed that both the absolute dissociation rate constant for T4 gene 32 protein and its salt dependence are sensitive to the average molecular weight and polydispersity of the nucleic acid sample used. This is a general phenomenon exhibited by proteins that bind to nucleic acids in a highly cooperative manner.  相似文献   

10.
We have studied the interactions of single-stranded polyribonucleotides with murine leukemia virus structural proteins p10, p10' (a p10 variant), and Pr65gag, as well as Rous sarcoma virus (RSV) pp12 (a p10 analog). Two quantitative assays have been used to monitor protein-RNA association: the fluorescence enhancement of polyethenoadenylic acid) poly(epsilon A) upon binding protein, and tryptophan fluorescence quenching upon binding to poly(U). With each assay p10 was shown to bind stoichiometrically to single-stranded RNA, covering a length of nucleic acid chain (occluded site size, n) of about 6 residues. RSV pp12 was also shown to bind to poly(epsilon A), with n = 5 +/- 1. Addition of NaCl to fully titrated MuLV p10-nucleic acid mixtures effected nearly complete restoration of poly(epsilon A) or MuLV p10 fluorescence. Under conditions of 0.06 M NaCl, p10 bound noncooperatively to poly(epsilon A) with an intrinsic association constant, K = 2.3 X 10(6) M-1. K and n determined in this study were shown to relate to Kapp determined by other methods, by the approximation Kapp approximately NK, where N is the number of binding sites along the polynucleotide chain ((nucleotides/chain)/n). Chemical modifications of the p10 cysteine residues did not alter the affinity for poly(epsilon A). The affinity of Pr65gag for poly(epsilon A) appears to be higher than that of p10.  相似文献   

11.
Binding of the single-stranded DNA-binding protein (SSB) of Escherichia coli to single-stranded (ss) polynucleotides produces characteristic changes in the absorbance (OD) and circular dichroism (CD) spectra of the polynucleotides. By use of these techniques, complexes of SSB protein and poly(rA) were shown to display two of the binding modes reported by Lohman and Overman [Lohman, T.M., & Overman, L. (1985) J. Biol. Chem. 260, 3594-3603]. The circular dichroism spectra of the "low salt" (10 mM NaCl) and "high salt" (greater than 50 mM NaCl) binding mode are similar in shape, but not in intensity. SSB binding to poly(rA) yields a complexed CD spectrum that shares several characteristics with the spectra obtained for the binding of AdDBP, GP32, and gene V protein to poly(rA). We therefore propose that the local structure of the SSB-poly(rA) complex is comparable to the structures proposed for the complexes of these three-stranded DNA-binding proteins with DNA (and RNA) and independent of the SSB-binding mode. Electric field induced birefringence experiments were used to show that the projected base-base distance of the complex is about 0.23 nm, in agreement with electron microscopy results. Nevertheless, the local distance between the successive bases in the complex will be quite large, due to the coiling of the DNA around the SSB tetramer, thus partly explaining the observed CD changes induced upon complexation with single-stranded DNA and RNA.  相似文献   

12.
Escherichia coli single-strand (ss) DNA binding protein (SSB) is an essential protein that binds ssDNA intermediates formed during genome maintenance. SSB homotetramers bind ssDNA in two major modes, differing in occluded site size and cooperativity. The (SSB)35 mode in which ssDNA wraps, on average, around two subunits is favored at low [NaCl] and high SSB/DNA ratios and displays high unlimited, nearest-neighbor cooperativity forming long protein clusters. The (SSB)65 mode, in which ssDNA wraps completely around four subunits of the tetramer, is favored at higher [NaCl] (>200 mM) and displays limited low cooperativity. Crystal structures of E. coli SSB and Plasmodium falciparum SSB show ssDNA bound to the SSB subunits (OB folds) with opposite polarities of the sugar phosphate backbones. To investigate whether SSB subunits show a polarity preference for binding ssDNA, we examined EcSSB and PfSSB binding to a series of (dT)70 constructs in which the backbone polarity was switched in the middle of the DNA by incorporating a reverse-polarity (RP) phosphodiester linkage, either 3′-3′ or 5′-5′. We find only minor effects on the DNA binding properties for these RP constructs, although (dT)70 with a 3′-3′ polarity switch shows decreased affinity for EcSSB in the (SSB)65 mode and lower cooperativity in the (SSB)35 mode. However, (dT)70 in which every phosphodiester linkage is reversed does not form a completely wrapped (SSB)65 mode but, rather, binds EcSSB in the (SSB)35 mode with little cooperativity. In contrast, PfSSB, which binds ssDNA only in an (SSB)65 mode and with opposite backbone polarity and different topology, shows little effect of backbone polarity on its DNA binding properties. We present structural models suggesting that strict backbone polarity can be maintained for ssDNA binding to the individual OB folds if there is a change in ssDNA wrapping topology of the RP ssDNA.  相似文献   

13.
The single-stranded DNA-binding proteins from bacteriophage T4, F plasmid, Escherichia coli, and calf thymus can all be covalently cross-linked in vitro to thymine oligonucleotides by irradiating the respective protein-oligonucleotide complexes with ultraviolet light. More extensive studies on the E. coli single-stranded DNA-binding protein (SSB) indicate that this reaction is dependent upon both the length of the oligonucleotide and the dose of ultraviolet irradiation. Using anion-exchange and reverse-phase ion-pairing high-performance liquid chromatography we have isolated a specific cross-linked tryptic peptide comprising residues 57-62 of the SSB protein with the sequence valine-valine-leucine-phenylalanine-glycine-lysine. Solid-phase sequence analysis of the covalent [32P] p(dT)8-peptide complex indicates that phenylalanine 60 is the site of cross-linking. This amino acid is located within the general region of SSB (residues 1-115) that has previously been shown to contain the DNA-binding site (Williams, K. R., Spicer, E. K., LoPresti, M. B., Guggenheimer, R. A., and Chase, J. W. (1983) J. Biol. Chem. 258, 3346-3355). The high-performance liquid chromatography purification procedure we have devised to isolate cross-linked peptide-oligonucleotide complexes should be of general applicability and should facilitate future structure/function studies on other nucleic acid-binding proteins.  相似文献   

14.
Kozlov AG  Lohman TM 《Biochemistry》2002,41(19):6032-6044
We have examined the kinetic mechanism for binding of the homotetrameric Escherichia coliSSB protein to single-stranded oligodeoxynucleotides [(dT)(70) and (dT)(35)] under conditions that favor the formation of a fully wrapped ssDNA complex in which all four subunits interact with DNA. Under these conditions, a so-called (SSB)(65) complex is formed in which either one molecule of (dT)(70) or two molecules of (dT)(35) bind per tetramer. Stopped-flow studies monitoring quenching of the intrinsic SSB Trp fluorescence were used to examine the initial binding step. To examine the kinetics of ssDNA wrapping, we used a single-stranded oligodeoxythymidylate, (dT)(66), that was labeled on its 3'-end with a fluorescent donor (Cy3) and on its 5'-end with a fluorescent acceptor (Cy5). Formation of the fully wrapped structure was accompanied by extensive fluorescence resonance energy transfer (FRET) from Cy3 to Cy5 since the two ends of (dT)(66) are in close proximity in the fully wrapped complex. Our results indicate that initial ssDNA binding to the tetramer is very rapid, with a bimolecular rate constant, k(1,app), of nearly 10(9) M(-1) s(-1) in the limit of low salt concentration (<0.2 M NaCl, pH 8.1, 25.0 degrees C), whereas the rate of dissociation is very low at all salt concentrations that were examined (20 mM to 2 M NaCl or NaBr). However, the rate of initial binding and the rate of formation of the fully wrapped complex are identical, indicating that the rate of wrapping of the ssDNA around the SSB tetramer is very rapid, with a lower limit rate of 700 s(-1). The implications of this rapid binding and wrapping reaction are discussed.  相似文献   

15.
Limited proteolysis can be used to remove either 42 or 62 amino acids at the COOH terminus of the 18,873-dalton Escherichia coli single-stranded DNA binding protein (SSB). Since poly(dT), but not d(pT)16, increases the rate of this reaction, it appears that cooperative SSB binding to single-stranded DNA (ssDNA) is associated with a conformational change that increases the exposure of the COOH terminus to proteolysis. As a result of this DNA-induced conformational change, we presume that the COOH-terminal region of SSB will become more accessible for interacting with other proteins that utilize the SSB:ssDNA complex as a substrate and that are involved in E. coli DNA replication, repair, and recombination. Removal of this COOH-terminal domain from SSB results in a stronger helix-destabilizing protein which suggests this region may be important for controlling the ability of SSB to denature double-stranded DNA. Since similar results have previously been reported for the bacteriophage T4 gene 32 protein (Williams, K.R., and Konigsberg, W. (1978) J. Biol. Chem. 253, 2463-2470; Hosoda, J., and Moise, H. (1978) J. Biol. Chem. 253, 7547-7555), the acidic, COOH-terminal domains of these two single-stranded DNA binding proteins may be functionally homologous. Preliminary evidence is cited that suggests other prokaryotic and eukaryotic DNA binding proteins may contain similar functional domains essential for controlling their ability to invade double helical DNA.  相似文献   

16.
This study reports on various parameters which affect the binding stoichiometry for complexes of bacteriophage T4 gene 32 protein (P32) and single stranded polynucleotides (determined by UV absorbance and fluorescence quenching) and presents results of a quantitative electron spin resonance assay to determine physiologically effective binding affinity differences of nucleic acid binding proteins. The assay employs macromolecular spin probes (spin-labeled nucleic acids) which are used to determine the fraction of saturation in competition experiments with unlabeled nucleic acids. It was found that the fraction of complexed spin-labeled polynucleotides can be directly monitored by ESR with a two-component analysis approach when ligands such as poly(L-lysine), gene 5 protein (P5) of filamentous bacteriophage fd, and gene 32 protein (P32) of bacteriophage T4 are used. The ESR data unequivocally show that: 1) the binding stoichiometry for poly(L-lysine), P5 and P32 is nucleotide/lysine, 4 nucleotides/P5 monomer, and 10 nucleotides/P32 monomer, respectively; and 2) under physiologically relevant buffer conditions the relative affinity of P32 in the cooperative binding mode for polythymidylic acid is about 4 times greater than for polydeoxyinosinic acid and about 12 times greater than for polyinosinic acid, and the relative affinity of P32 for polydeoxyinosinic acid is about 3 times greater than for polyinosinic acid.  相似文献   

17.
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   

18.
Many macromolecular interactions, including protein‐nucleic acid interactions, are accompanied by a substantial negative heat capacity change, the molecular origins of which have generated substantial interest. We have shown previously that temperature‐dependent unstacking of the bases within oligo(dA) upon binding to the Escherichia coli SSB tetramer dominates the binding enthalpy, ΔHobs, and accounts for as much as a half of the observed heat capacity change, ΔCp. However, there is still a substantial ΔCp associated with SSB binding to ssDNA, such as oligo(dT), that does not undergo substantial base stacking. In an attempt to determine the origins of this heat capacity change, we have examined by isothermal titration calorimetry (ITC) the equilibrium binding of dT(pT)34 to SSB over a broad pH range (pH 5.0–10.0) at 0.02 M, 0.2 M NaCl and 1 M NaCl (25°C), and as a function of temperature at pH 8.1. A net protonation of the SSB protein occurs upon dT(pT)34 binding over this entire pH range, with contributions from at least three sets of protonation sites (pKa1 = 5.9–6.6, pKa2 = 8.2–8.4, and pKa3 = 10.2–10.3) and these protonation equilibria contribute substantially to the observed ΔH and ΔCp for the SSB‐dT(pT)34 interaction. The contribution of this coupled protonation (∼ −260 to −320 cal mol−1 K−1) accounts for as much as half of the total ΔCp. The values of the “intrinsic” ΔCp,0 range from −210 ± 33 cal mol−1 °K−1 to −237 ± 36 cal mol−1K−1, independent of [NaCl]. These results indicate that the coupling of a temperature‐dependent protonation equilibria to a macromolecular interaction can result in a large negative ΔCp, and this finding needs to be considered in interpretations of the molecular origins of heat capacity changes associated with ligand‐macromolecular interactions, as well as protein folding. Proteins 2000;Suppl 4:8–22. © 2000 Wiley‐Liss, Inc.  相似文献   

19.
The bacteriophage T4 regA protein (M(r) = 14,6000) is a translational repressor of a group of T4 early mRNAs. To identify a domain of regA protein that is involved in nucleic acid binding, ultraviolet light was used to photochemically cross-link regA protein to [32P]p(dT)16. The cross-linked complex was subsequently digested with trypsin, and peptides were purified using anion exchange high performance liquid chromatography. Two tryptic peptides cross-linked to [32P]p(dT)16 were isolated. Gas-phase sequencing of the major cross-linked peptide yielded the following sequence: VISXKQKHEWK, which corresponds to residues 103-113 of regA protein. Phenylalanine 106 was identified as the site of cross-linking, thus placing this residue at the interface of the regA protein-p(dT)16 complex. The minor cross-linked peptide corresponded to residues 31-41, and the site of cross-linking in the peptide was tentatively assigned to Cys-36. The nucleic acid binding domain of regA protein was further examined by chemical cleavage of regA protein into six peptides using CNBr. Peptide CN6, which extends from residue 95 to 122, retains both the ability to be cross-linked to [32P]p(dT)16 and 70% of the nonspecific binding energy of the intact protein. However, peptide CN6 does not exhibit the binding specificity of the intact protein. Three of the other individual CNBr peptides have no measurable affinity for nucleic acid, as assayed by photo-cross-linking or gel mobility shifts.  相似文献   

20.
The effect of Mg2+ on the binding of the Streptococcus pneumoniae single-stranded DNA binding (SSB) proteins, SsbA and SsbB, to various dT(n) oligomers was examined by polyacrylamide gel electrophoresis. The results were then compared with those that were obtained with the well characterized SSB protein from Escherichia coli, SsbEc. In the absence of Mg2+, the results indicated that the SsbEc protein was able to bind to the dT(n) oligomers in the SSB(35) mode, with only two of the four subunits of the tetramer interacting with the dT(n) oligomers. In the presence of Mg2+, however, the results indicated that the SsbEc protein was bound to the dT(n) oligomers in the SSB(65) mode, with all four subunits of the tetramer interacting with the dT(n) oligomers. The SsbA protein behaved similarly to the SsbEc protein under all conditions, indicating that it undergoes Mg2+ -dependent changes in its DNA binding modes that are analogous to those of the SsbEc protein. The SsbB protein, in contrast, appeared to bind to the dT(n) oligomers in an SSB(65)-like mode in either the presence or the absence of Mg2+, suggesting that it may not exhibit the pronounced negative intrasubunit cooperativity in the absence of Mg2+ that is required for the formation of the SSB(35) mode. Additional experiments with a chimeric SsbA/B protein indicated that the structural determinants that govern the transitions between the different DNA binding modes may be contained within the N-terminal domains of the SSB proteins.  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