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.
The influence of additional operator or pseudooperator sequences on the lactose repressor-operator interaction has been investigated. Results of kinetic and equilibrium binding measurements suggest an important in vivo role for the Z-gene pseudooperator in repressor-operator binding; the formation of a ternary, looped complex is indicated by the influence of secondary operator sites on binding parameters. Although the binding affinity of the Z-gene pseudooperator [Oz] is only approximately 1/30 that observed for the primary operator [O], the binding affinity to DNA containing both Oz and O is significantly higher than either sequence alone or the sum of the two. This synergistic effect is enhanced further by replacing the pseudooperator sequence [Oz] with the primary operator sequence and results in an even stronger ternary complex in plasmids with duplicate primary sites. The distance between the center of the two primary operators affects the formation of a ternary complex in the linear DNA molecules. Decreased dissociation rate constants were observed with spacing of operator-like sequences between 300 and 500 base pairs (bp). Minimal influence of the second operator on repressor binding is observed when the operators are separated by approximately 4000 or approximately 100 bp. The significant influence of distance on kinetic and equilibrium parameters was demonstrated by measurements on plasmid pRW1511 [Oi-O-PL-Oz] cleaved with restriction enzymes either in the polylinker region to place Oi-O and Oz on opposite ends of the linear plasmid or outside this region to maintain the sites within 500 bp. These results are consistent with the formation of operator-repressor-pseudooperator ternary complex to generate a looped DNA structure.  相似文献   

2.
3.
P A Whitson  K S Matthews 《Biochemistry》1986,25(13):3845-3852
The dissociation kinetics for repressor-32P-labeled operator DNA have been examined by adding unlabeled operator DNA to trap released repressor or by adding a small volume of concentrated salt solution to shift the Kd of repressor-operator interaction. The dissociation rate constant for pLA 322-8, an operator-containing derivative of pBR 322, was 2.4 X 10(-3) s-1 in 0.15 M KCl. The dissociation rate constant at 0.15 M KCl for both lambda plac and pIQ, each of which contain two pseudooperator sequences, was approximately 6 X 10(-4) s-1. Elimination of flanking nonspecific DNA sequences by use of a 40 base pair operator-containing DNA fragment yielded a dissociation rate constant of 9.3 X 10(-3) s-1. The size and salt dependences of the rate constants suggest that dissociation occurs as a multistep process. The data for all the DNAs examined are consistent with a sliding mechanism of facilitated diffusion to/from the operator site. The ability to form a ternary complex of two operators per repressor, determined by stoichiometry measurements, and the diminished dissociation rates in the presence of intramolecular nonspecific and pseudooperator DNA sites suggest the formation of an intramolecular ternary complex. The salt dependence of the dissociation rate constant for pLA 322-8 at high salt concentrations converges with that for a 40 base pair operator. The similarity in dissociation rate constants for pLA 322-8 and a 40 base pair operator fragment under these conditions indicates a common dissociation mechanism from a primary operator site on the repressor.  相似文献   

4.
Thermodynamic analysis of the lactose repressor-operator DNA interaction   总被引:4,自引:0,他引:4  
Kinetic and equilibrium constants for lactose repressor-operator DNA interaction have been examined as a function of salt concentration, size and sequence context of the operator DNA, and temperature. Significant salt effects were observed on kinetic and equilibrium parameters for pLA 322-8, an operator-containing derivative of pBR 322, and pIQ, an operator and pseudooperator-containing derivative of pBR 322. The association rate constant and equilibrium constant for the 40 base pair operator fragment were also salt dependent. Data for all the DNAs were consistent with a sliding mechanism for repressor-operator association/dissociation [Berg, O. G., & Blomberg, C. (1978) Biophys. Chem. 8, 271-280]. Calculation of the number of ionic interactions based on salt dependence yielded a value of approximately 8 for repressor binding to pIQ and pLA 322-8 vs. approximately 6 for the repressor-40 base pair fragment. These data and the differences in binding parameters for the plasmids vs. the 40 base pair operator are consistent with the formation of an intramolecular ternary complex in the plasmid DNAs. Unusual biphasic temperature dependence was observed in the equilibrium and dissociation rate constants for pLA 322-8, pIQ, and the 40 base pair fragment. These observations coupled with a discontinuity found in the inducer association rate constant as a function of temperature suggest a structural change in the protein. The large positive entropy contributions associated with repressor binding to all the DNAs examined provide the significant driving force for the reaction and are consistent with involvement of ionic and apolar interactions in complex formation.  相似文献   

5.
The equilibrium association constants for the binding of a wide variety of effecting ligands of the lac repressor were measured by equilibrium dialysis. Also, detailed investigations of the apparent rate of dissociation of repressor-operator comples as a function of ligand concentration were carried out for several inducers and anti-inducers. The affinity of repressor-ligand comples for operator DNA was evaluated from the specific rate constants at saturating concentrations of effecting ligand. By fitting the experimental data depicting the functional dependence of the rate of dissociation upon ligand concentrations to calculated curves, assuming simple models of the induction mechanism, the equilibrium association constant for the binding of effecting ligand to repressor-operator comples was determined. Inducers reduce the affinity of lac repressor for operator DNA by a factor of approximately 1000 under standard conditions; the extent of destabilization depends on Mg2+ ion concentration. Anti-inducers increase the affinity of repressor for operator at most a factor of five. Only one neutral ligand, which binds to repressor without altering the stability of repressor-operator comples, was found. No homotropic or heterotropic interactions in the binding of effecting ligands either to repressor or to repressor-operator complex are evident.  相似文献   

6.
Analysis of trp repressor-operator interaction by filter binding.   总被引:6,自引:1,他引:5       下载免费PDF全文
A filter binding assay was developed that allows measurement of specific binding of trp repressor to operator DNA. The most important feature of this procedure is the concentration and type of salt present in the binding buffer. Using this assay the dissociation constant of the repressor-operator complex was determined to be 2.6 X 10(-9) M, and 1.34 repressor dimers were found to be bound to each operator-containing DNA molecule. These values agree with those obtained by more complex methods. The dissociation constant of the repressor for the corepressor L-tryptophan in the presence of operator DNA was shown to be 2.5 X 10(-5) M. A synthetic 48 bp operator fragment was used to determine the repressor-operator dissociation constant in the presence of tryptophan or tryptophan analogs which have higher or lower affinities for aporepressor. The rate of dissociation of repressor from operator DNA also was determined. Our findings indicate that dissociation is influenced by the concentration of tryptophan or tryptophan analogs and suggest that release of the corepressor may be the first step in dissociation of the repressor-operator complex.  相似文献   

7.
The nitrocellulose filter assay was used to study the effect of the DNA denaturants glycerol and dimethylsulfoxide (Me2SO) on the lac repressor-operator interaction. Both glycerol and Me2SO decrease the rate of dissociation (kb) of the repressor-operator complex but do not significantly alter the rate of association of repressor and operator. In the presence of 10% Me2SO an almost 10-fold increase of affinity of repressor for operator is observed. A small increase in affinity of repressor for Escherichia coli DNA, chicken blood DNA, and poly(dA-dT) is also found. The results lead to the conclusion that lac repressor when interacting with the operator causes local destabilization of the DNA.  相似文献   

8.
The TOP3 gene of the yeast Saccharomyces cerevisiae was postulated to encode a DNA topoisomerase, based on its sequence homology to Escherichia coli DNA topoisomerase I and the suppression of the poor growth phenotype of top3 mutants by the expression of the E. coli enzyme (Wallis, J.W., Chrebet, G., Brodsky, G., Golfe, M., and Rothstein, R. (1989) Cell 58, 409-419). We have purified the yeast TOP3 gene product to near homogeneity as a 74-kDA protein from yeast cells lacking DNA topoisomerase I and overexpressing a plasmid-borne TOP3 gene linked to a phosphate-regulated yeast PHO5 gene promoter. The purified protein possesses a distinct DNA topoisomerase activity: similar to E. coli DNA topoisomerases I and III, it partially relaxes negatively but not positively supercoiled DNA. Several experiments, including the use of a negatively supercoiled heteroduplex DNA containing a 29-nucleotide single-stranded loop, indicate that the activity has a strong preference for single-stranded DNA. A protein-DNA covalent complex in which the 74-kDa protein is linked to a 5' DNA phosphoryl group has been identified, and the nucleotide sequences of 30 sites of DNA-protein covalent complex formation have been determined. These sequences differ from those recognized by E. coli DNA topoisomerase I but resemble those recognized by E. coli DNA topoisomerase III. Based on these results, the yeast TOP3 gene product can formally be termed S. cerevisiae DNA topoisomerase III. Analysis of supercoiling of intracellular yeast plasmids in various DNA topoisomerase mutants indicates that yeast DNA topoisomerase III has at most a weak activity in relaxing negatively supercoiled double-stranded DNA in vivo, in accordance with the characteristics of the purified enzyme.  相似文献   

9.
A method has been used to quantitate the reaction between eukaryotic type I DNA topoisomerase and topological forms of DNA. This procedure (Trask, D.K., DiDonato, J.D. and Muller, M.T. (1984) Eur. Mol. Biol. Organ. J. 3, 671-676) measures the efficiency of DNA cleavage and concurrent formation of a covalent enzyme/DNA complex. Eukaryotic type I topoisomerases react preferentially by 5-10-fold with supercoiled DNA. The effect of supercoiling is clearly evident in that both the initial rate and final extent of the reaction is elevated. Because the dissociation rate is much lower than the association rate, it is possible to isolate native topoisomerase/DNA complexes. These complexes are comprised of enzyme molecules which are catalytically active when challenged with a second supercoiled DNA substrate. Collectively, the data support the conclusion that a functional intermediate in the reaction sequence is being detected and that the avian topoisomerase I preferentially cleaves supercoiled DNA.  相似文献   

10.
Equilibrium and kinetic studies on the folding of a series of amino acid replacements at position 211 in the alpha subunit of tryptophan synthase from Escherichia coli were performed in order to determine the role of this position in the rate-limiting step in folding. Previous studies [Beasty, A. M., Hurle, M. R., Manz, J. T., Stackhouse, T., Onuffer, J. J., & Matthews, C. R. (1986) Biochemistry 25, 2965-2974] have shown that the rate-limiting step corresponds to the association/dissociation of the amino (residues 1-188) and carboxy (residues 189-268) folding units. In terms of the secondary structure, the amino folding unit consists of the first six strands and five alpha helices of this alpha/beta barrel protein. The carboxy folding unit comprises the remaining two strands and three alpha helices; position 211 is in strand 7. Replacement of the wild-type glycine at position 211 with serine, valine, and tryptophan at most alters the rate of dissociation of the folding units; association is not changed significantly. In contrast, glutamic acid and arginine dramatically decelerate and accelerate, respectively, both association and dissociation. The difference in effects is attributed to long-range electrostatic interactions for these charged side chains; steric effects and/or hydrogen bonding play lesser roles. When considered with previous data on replacements at other positions in the alpha subunit [Hurle, M. R., Tweedy, N. B., & Matthews, C. R. (1986) Biochemistry 25, 6356-6360], it is clear that beta strands 6 (in the amino folding unit) and 7 (in the carboxy folding unit and containing position 211) dock late in the folding process.  相似文献   

11.
The carboxyl-terminal sequence of the lac repressor protein contains heptad repeats of leucines at positions 342, 349, and 356 that are required for tetramer assembly, as substitution of these leucine residues yields solely dimeric species (Chakerian, A. E., Tesmer, V. M., Manly, S. P., Brackett, J. K., Lynch, M. J., Hoh, J. T., and Matthews, K. S. (1991) J. Biol. Chem. 266, 1371-1374; Alberti, S., Oehler, S., von Wilcken-Bergmann, B., Kr?mer, H., and Müller-Hill, B. (1991) New Biol. 3, 57-62). To further investigate this region, which may form a leucine zipper motif, a family of lac repressor carboxyl-terminal deletion mutants eliminating the last 4, 5, 11, 18, and 32 amino acids (aa) has been constructed. The -4 aa mutant, in which all of the leucines in the presumed leucine zipper are intact, is tetrameric and displays operator and inducer binding properties similar to wild-type repressor. The -5 aa, -11 aa, -18 aa, and -32 aa deletion mutants, depleted of 1, 2, or all 3 of the leucines in the heptad repeats, are all dimeric, as demonstrated by gel filtration chromatography. Circular dichroism spectra and protease digestion studies indicate similar secondary/tertiary structures for the mutant and wild-type proteins. Differences in reaction with a monoclonal antibody specific for a subunit interface are observed for the dimeric versus tetrameric proteins, indicative of exposure of the target epitope as a consequence of deletion. Inducer binding properties of the deletion mutants are similar to wild-type tetrameric repressor at neutral pH. Only small differences in affinity and cooperativity from wild-type are evident at elevated pH; thus, the cooperative unit within the tetramer appears to be the dimer. "Apparent" operator binding affinity for the dimeric proteins is diminished, although minimal change in operator dissociation rate constants was observed. The diminution in apparent operator affinity may therefore derive from either 1) dissociation of the dimeric mutants to monomer generating a linked equilibrium or 2) alterations in intrinsic operator affinity of the dimers; the former explanation is favored. This detailed characterization of the purified mutant proteins confirms that the carboxyl-terminal region is involved in the dimer-dimer interface and demonstrates that cooperativity for inducer binding is contained within the dimer unit of the tetramer structure.  相似文献   

12.
M E Dahlberg  S J Benkovic 《Biochemistry》1991,30(20):4835-4843
In a previously determined minimal kinetic scheme for DNA polymerization catalyzed by the Klenow fragment (KF) of Escherichia coli DNA polymerase I, a nonchemical step that interconverted the KF'.DNAn+1.PPi and KF.DNAn+1PPi complexes was not observed in correct incorporation [Kuchta, R. D., Mizrahi, V., Benkovic, P.A., Johnson, K.A., & Benkovic, S.J. (1987) Biochemistry 26, 8410-8417] but was detected in misincorporation [Kuchta, R. D., Benkovic, P.A., & Benkovic, S.J. (1988) Biochemistry 27, 6716-6725]. In a pulse-chase experiment in this study, a burst amplitude of 100% of the enzyme concentration is observed; under pulse-quench conditions, the burst amplitude is 80%, indicative of the accumulation of the KF'.DNA.dNTP species owing to a slow step subsequent to chemical bond formation. This latter step was unequivocally identified by single-turnover pyrophosphorolysis and pyrophosphate-exchange experiments as one interconverting KF'.DNAn+1.PPi and KF.DNAn+1.PPi. The rate constants for this step in both directions were established through the rate constants for processive synthesis and pyrophosphorolysis. Pyrophosphorolysis of a 3'-phosphorothioate DNA duplex confirmed that the large elemental effect observed previously [Mizrahi, V., Henrie, R. N., Marlier, J.F., Johnson, K.A., & Benkovic, S.J. (1985) Biochemistry 24, 4010-4018] in this direction but not in polymerization is due to a marked decrease in the affinity of KF for the phosphorothioate-substituted duplex and not to the chemical step. The combination of the experimentally measured equilibrium constant for the bound KF.DNA species with the collective kinetic measurements further extends previous insights into the dynamics of the polymerization process catalyzed by KF.  相似文献   

13.
14.
The kinetics of ATP binding and hydrolysis (formation of acid-labile phosphate) by the Tetrahymena 30 S dynein ATPase has been measured by chemical quench flow methods. The amplitude of the ATP-binding transient gave a molecular weight per ATP-binding site of approximately 750,000, suggesting nearly 3 ATP binding sites/2 million Mr dynein molecule (Johnson, K. A., and Wall, J.S. (1983) J. Cell Biol. 96, 669-678). ATP binding occurred at the rate predicted from the apparent second order rate constant of 4.7 X 10(6) M-1 S-1 measured by analysis of the ATP-induced dissociation of the microtubule-dynein complex (Porter, M. E., and Johnson, K. A. (1983) J. Biol. Chem. 258, 6582-6587). Hydrolysis was slower than binding and occurred at a rate of 55 S-1, at 30 and 50 microM ATP. The rate limiting step for steady state turnover (product release) occurred with a rate constant of 8 S-1. These data show that the first two steps of the pathway of coupling ATP hydrolysis to the microtubule-dynein cross-bridge cycle are the same as those described by Lymn and Taylor for actomyosin (Lymn, R. W., and Taylor, E. W. (1971) Biochemistry 10, 4617-4624). Namely, ATP binding induces the very rapid dissociation of dynein from the microtubule and ATP hydrolysis occurs more slowly following dissociation. Moreover, in spite of rather gross structural differences, the kinetic constants for dynein and myosin are quite similar.  相似文献   

15.
Molecular modeling studies [Islam, S.A., Neidle, S., Gandecha, B.M., Partridge, M., Patterson, L.H., & Brown, J.R. (1985) J. Med. Chem. 28, 857-864] have suggested that anthracene-9,10-dione (anthraquinone) derivatives substituted at the 1,4 and 1,8 positions with-NH(CH2)2NH(CH2CH3)2+ side chains intercalate with DNA with both substituents in the same groove (classical intercalation) while a similarly substituted 1,5 derivative intercalates in a threading mode with one side chain in each groove. Modeling studies also suggested that anthracene-9,10-dione (anthraquinone) derivatives substituted at the 2,6 positions with -NHCO(CH2)R (where R is a cationic group) should bind to DNA by the threading mode, and several such derivatives have been synthesized [Agbandjie, M., Jenkins, T.C., McKenna, R., Reszka, A., & Neidle, S. (1992) J. Med. Chem. 35, 1418-1429]. We have conducted stopped-flow kinetics association and dissociation experiments on the interaction of these anthraquinones with calf thymus DNA and with DNA polymers with alternating AT and GC base pairs to experimentally determine the binding mode and how the threading mode affects intercalation rates relative to similarly substituted classical intercalators. The binding modes, determined by analysis of relative rates, energies of activation, and effects of salt concentration on association and dissociation rate constants, agree completely with the modes predicted by molecular modeling methods. Association and dissociation rate constants for the threading mode are approximately a factor of 10 lower than constants for the classical intercalation mode, and the two modes, thus, have similar binding constants. Variations in rate constants for changes in cationic substituents at the 2 and 6 positions of the anthraquinone ring were surprisingly small.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
17.
Proteins that recognize and bind specific sites in DNA are essential for regulation of numerous biological functions. Such proteins often require a negative supercoiled DNA topology to function correctly. In current research, short linear DNA is often used to study DNA–protein interactions. Although linear DNA can easily be modified, for capture on a surface, its relaxed topology does not accurately resemble the natural situation in which DNA is generally negatively supercoiled. Moreover, specific binding sequences are flanked by large stretches of non-target sequence in vivo. Here, we present a straightforward method for capturing negatively supercoiled plasmid DNA on a streptavidin surface. It relies on the formation of a temporary parallel triplex, using a triple helix forming oligonucleotide containing locked nucleic acid nucleotides. All materials required for this method are commercially available. Lac repressor binding to its operator was used as model system. Although the dissociation constants for both the linear and plasmid-based operator are in the range of 4 nM, the association and dissociation rates of Lac repressor binding to the plasmid-based operator are ∼18 times slower than on a linear fragment. This difference underscores the importance of using a physiologically relevant DNA topology for studying DNA–protein interactions.  相似文献   

18.
19.
We have analyzed protein-DNA complexes formed between lac repressor and linear or differently supercoiled lac DNA (802 or 816 base-pairs in length), which carry all three natural lac operators (O1, O2 and O3) in their wild-type sequence context and spacing and compared them with constructs that contain specifically mutated "pseudo-operators" O2 or O3. We used gel retardation assays to identify the nature of the complexes according to their characteristic electrophoretic mobility and dissociation rate measurements to determine their stability. With linear DNA we found only indirect evidence for loop formation between O1 and O2. In covalently closed DNA minicircles the formation of a loop between O1 and O2 could be demonstrated by the observation that O1-O2 containing DNA with low negative supercoiling (sigma = -0.013 and less) is constricted by binding of lac repressor, resulting in an increased electrophoretic mobility. At elevated negative supercoiling (sigma = -0.025, -0.037, -0.05) O1-O2 containing DNA complexed with lac repressor migrates significantly slower than the corresponding O1-DNA, indicating loop formation. The dissociation of lac repressor-operator complexes is decreased with increasing negative supercoiling for all tested operator combinations of O1, O2 and O3. However, in the presence of at least two natural lac operators on the same DNA minicircle the enhancement of stability is particularly large. This indicates that a DNA loop is formed between these two lac operators, O1 and O2 as well as O1 and O3, since negative supercoiling is known specifically to promote the formation of looped structures. Additionally, we observe a dependence of dissociation rate on the spatial alignment of the operators as a result of changing helical periodicity in differently supercoiled DNA and consider this to be further evidence for loop formation between O1 and O2 as well as O1 and O3.  相似文献   

20.
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