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.
In vitro measurements show that the X86 repressor, which has an increased affinity for the lac operator as compared to wild-type repressor, also has an increased affinity for non-operator sites on Escherichia coli DNA. The rate constant of association of repressor and operator is decreased by E. coli DNA fivefold more for X86 repressor than for wild-type repressor. Low inducer concentrations increase the rate of association of X86 repressor and operator in the presence of E. coli DNA. In a partial equilibrium situation where part of the X86 repressor is bound to the operator, and part to either non-operator sites on E. coli DNA or to an Oc operator, the formation of complexes between X86 repressor and wild-type operator is favored by low inducer concentrations. Repression of the lac enzymes increases drastically in the X86 mutant in the absence of DNA synthesis in vivo. A new explanation for the in vivo characteristics of the X86 mutant is suggested.  相似文献   

2.
Using the protein predictive model of Chou & Fasman (1974b), the secondary structure of the lac repressor has been elucidated from its amino acid sequence of 347 residues. The conformation is predicted to contain 37% α-helix and 35% β-sheet for the repressor, and 29% helix and 41% β-sheet for the trypsin-resistant core (residues 60 to 327). Circular dichroism studies indicate that native lac repressor contains 40% helix and 42% β-sheet, while the core has 16% helix and 54% β-sheet, in general agreement with the predicted conformation. The sharp reduction in helicity for the trypsinized lac repressor could be due to the loss of two long helical regions, 26–45 and 328–344, predicted at both terminals. There are extensive β-sheets predicted in the 215–324 region, which may be responsible for tetrameric stabilization found in both the lac repressor and the core. Residues 17 to 33 were previously predicted by Adler et al. (1972) to be helical and were proposed to bind in the major groove of DNA. However, the present analysis shows that there are two anti-parallel β-sheet regions: 4–7 and 17–24 at the N-terminal as well as 315–318 and 321–324 at the C-terminal of the lac repressor. These β-sheet pairs may assume the twisted “polypeptide double helix” conformation (Carter & Kraut, 1974) and bind to complementary regions in the major groove of DNA. The OH groups of Tyr at the N-terminal and those of Thr and Ser side chains, in both β-sheets at the N and C-terminal ends, could form hydrogen bonds to specific sites on the lac operator. There are 23 reverse β-turns predicted that may control the tertiary folding of the lac repressor, which is essential for operator binding. The behavior of several lac repressor mutants can be satisfactorily explained in terms of polar to non-polar group replacements as well as conformational changes in light of the present predicted model.  相似文献   

3.
4.
Hypoxanthine was substituted for guanine at specific sites in the lac operator DNA by a combination of chemical and enzymatic procedures. The stability of these modified lac operators with wild-type (SQ) and tight binding (QX86) lac repressors was measured. Effects were variable. At some sites insertion of hypoxanthine significantly reduced the stability of the complex whereas at other sites substitution with hypoxanthine did not alter the repressor—operator interaction. In addition, insertion of this analog at two sites increased the stability of the complexes. These changes were used to partially map regions of the lac operator that are in contact with lac represser. The results suggest that lac repressor recognizes the guanine 2-amino group at specific sites in the minor groove of lac operator.  相似文献   

5.
A 203 base-pair fragment containing the lac operator/promoter region of Escherichia coli was inserted into the EcoRI site of the plasmid vector pKC7. Rates of restriction endonuclease cleavage of the flanking EcoRI sites and of several other restriction sites on the DNA molecule were then compared in the presence and absence of bound RNA polymerase or lac repressor. The rates were identical whether or not protein had been bound, even for sites as close as 40 base-pairs from a protein binding site. No difference was detected using supercoiled, nicked circular, or linear DNA substrates. No apparent change in the rates of methylation of EcoRI sites by EcoRI methylase was produced by binding the regulatory proteins.  相似文献   

6.
A genetic mapping system is used to locate mutations on the lac repressor gene (I) which lead to repressor proteins with an increased affinity for operator DNA. These tight binding repressors (Itb) are of particular interest since their analysis should allow some conclusions on the mechanism of interaction between repressor and operator. Itb mutations were found to map in two regions of the I gene. One is near the amino-terminal end, a region which has been shown to be essential for the DNA binding properties of the repressor. The other region in which Itb mutations were mapped codes for approximately amino acids 255 to 295 of the repressor, a region which had so far not been considered to be essential for the DNA binding properties of the repressor protein.  相似文献   

7.
Summary In the lac operon, the existence of a secondary repressor binding site, inside Z gene, had been inferred from in vitro binding studies (Reznikoff et al., 1974; Gilbert et al., 1975).A serie of deletions have been constructed from a lac transducing bacteriophage. Some of those deleted bacteriophages have still the property of derepressing a chromosomal lac operon, even though they do not contain any more the lac operator. This phenomenon is an indication that the secondary repressor binding site is also active in vivo.  相似文献   

8.
In order to compare the structures of the DNA-binding sites on variants of the lac repressor, we have studied the influence of these variants on the dimethylsulfate methylation of the lac operator. Since a bound protein changes the availability of specific purines in the operator to this chemical attack, comparisons of the methylation patterns will show similarities or differences in the protein DNA contacts. We compared lac repressor, induced lac repressor (repressor bound to the gratuitous inducer isopropyl-β-d-thiogalactoside), mutant repressors having increased operator affinities (X86, I12 and the X86-I12 double mutant) and repressor peptides (long headpiece, residues 1 to 59 and short headpiece. residues 1 to 51). All of these repressors and repressor peptides exhibit the same general pattern of protection and enhancement in the operator; however, the short headpiece pattern differs most from that of the repressor while the induced repressor and the long headpiece show intermediate patterns that are strikingly similar to each other. The mutant repressors do not show an isopropyl-β-d-thiogalactoside effect but otherwise are almost indistinguishable from wild-type repressor. These results demonstrate that all molecules bind to the operator using basically the same protein-DNA contacts; they imply that (1) most and possibly all repressor contacts to operator lie within amino acids 1 to 51, (2) inducer weakens many contacts rather than totally disrupting one or even a few and (3) the tight-binding mutants do not make additional contacts to the DNA.These results are consistent with a model in which the amino-terminal portions of two repressor monomers make the DNA contacts. We show that one can understand the affinity of binding as related to the accuracy of the register of the two amino-terminal portions along the DNA. Furthermore, the action of inducer and the behaviour of the tight binding mutants can be accomodated within a two-state model in which the strongly or weakly binding states correspond to structures in which the amino-terminal regions are rigidly or loosely held with respect to each other.  相似文献   

9.
Nucleotide sequences in two wild-type and six mutant operators in the DNA of phage λ are compared. Strikingly similar 17 base pair units are found which we identify as the repressor binding sites. Each operator contains multiple repressor binding sites separated by A-T rich spacers. Elements of 2 fold rotational symmetry are present in each of the sites. Superimposed on each operator is an E. coli RNA polymerase recognition site (promoter). Similarities in the sequences of the two λ promoters, a lac promoter, and an E. coli RNA polymerase recognition site in SV40 DNA are noted.  相似文献   

10.
The Escherichia coli lactose (lac) operon encodes the first genetic switch to be discovered, and lac remains a paradigm for studying negative and positive control of gene expression. Negative control is believed to involve competition of RNA polymerase and Lac repressor for overlapping binding sites. Contributions to the local Lac repressor concentration come from free repressor and repressor delivered to the operator from remote auxiliary operators by DNA looping. Long-standing questions persist concerning the actual role of DNA looping in the mechanism of promoter repression. Here, we use experiments in living bacteria to resolve four of these questions. We show that the distance dependence of repression enhancement is comparable for upstream and downstream auxiliary operators, confirming the hypothesis that repressor concentration increase is the principal mechanism of repression loops. We find that as few as four turns of DNA can be constrained in a stable loop by Lac repressor. We show that RNA polymerase is not trapped at repressed promoters. Finally, we show that constraining a promoter in a tight DNA loop is sufficient for repression even when promoter and operator do not overlap.  相似文献   

11.
Chen et al. have proved conclusively that lac repressor and RNA polymerase bind independently to wild type lac DNA in vitro. To explain the lacp s mutation, which causes competitive binding between repressor and polymerase, they suggest that a new promoter site has been created near the lac operator.  相似文献   

12.
The interaction oflacoperator DNA withlacrepressor (LacI) is a classic example of a genetic regulatory switch. To dissect the role of stoichiometry, subunit association, and effects of DNA length in positioning this switch, we have determined binding isotherms for the interaction of LacI with a high affinity (Osym) operator on linearized plasmid (2500 bp) DNA over a wide range of macromolecular concentrations (10−14to 10−8M). Binding data were analyzed using a thermodynamic model involving four equilibria: dissociation of tetramers (T) into dimers (D), and binding of operator-containing plasmid DNA (O) to dimers and tetramers to form three distinct complexes, DO, TO, and TO2. Over the range of con- centrations of repressor, operator, and salt (0.075 M K+to 0.40 M K+) investigated, we find no evidence for any significant thermodynamic effect of LacI dimers. Instead, all isotherms can be interpreted in terms of just two equilibria, involving only T and the TO and TO2complexes. As a reference binding equilibrium, which we propose must approximate the DO binding interaction, we compare the plasmid Osymresults with our extensive studies of the binding of a 40 bp OsymDNA fragment to LacI. On this basis, we obtain a lower bound on the LacI dimer – tetramer equilibrium constant and values of the equilibrium constants for formation of TO and TO2complexes.At a salt concentration of 0.40 M, the Osymplasmid binding data are consistent with a model with two independent and identical binding sites for operator per LacI tetramer, in which the binding to a site on the tetramer is only slightly more favorable than the reference binding interaction. Increasingly large deviations from the independent-site model are observed as the salt concentration is reduced; binding of a second operator to form TO2becomes strongly disfavored relative to formation of TO at low salt concentrations (0.075 to 0.125 M). In addition, binding of both the first and second plasmid operator DNA molecules to the tetramer becomes increasingly more favorable than the reference binding interaction as [K+] is reduced from 0.40 M to 0.125 M. At 0.075 M K+, however, the strength of binding of the second plasmid operator DNA to the LacI tetramer is dramatically reduced; this interaction is much less favorable than binding the first plasmid operator DNA, and becomes much less favorable than the reference binding interaction. We propose that these differences arise from changes in the nature of the TO and TO2complexes with decreasing salt concentration. At low salt concentration, we suggest the hypothesis that flanking non-operator sequences bind non-specifically (coulombically) by local wrapping, and that distant regions of non-operator DNA occupy the second operator-binding site by looping. We propose that wrapping stabilizes both 1:1 and 2:1 complexes at low salt concentration, and that looping stabilizes the 1:1 complex but competitively destabilizes the 2:1 TO2complex at low salt concentration. These effects must play a role in adjusting the stability and structure of the LacI-lac operator repression complex as the cytoplasmic [K+] varies in response to changes in extracellular osmolarity.  相似文献   

13.
We have examined the N-terminal 56 amino acid fragment, the domain that can bind DNA independently, from 3-fluorotyrosine-substituted Escherichia coli lac repressor by 19F-nuclear magnetic resonance. The fragments or “headpieces” from four altered repressers missing each of the tyrosines in turn were examined in parallel. When the wild-type N-terminal fragment is titrated with a 36 base-pair lac operator DNA sequence, the 19F resonances undergo changes in their chemical shifts that are different from those changes when the N-terminal fragment is titrated with non-specific DNA fragments. By looking at these operator-induced changes as well as pH-dependent effects with all four altered N-terminal fragments, we show systematic correlations with the genetic data. The data lead us to conclude that upon operator DNA binding: (1) tyrosine 7 is displaced to a less polar environment and the higher than normal pK value of the phenolic OH group is decreased; (2) tyrosine 12 does not change much in either its mobility or environment; and (3) tyrosine 17 is involved, as suggested by the genetic data, when the headpiece forms a complex with operator DNA.  相似文献   

14.
Complexes of Escherichia coli lac-repressor with non-operator DNA have been visualized in the electron microscope using high-resolution metal shadowing and negative staining. Under conditions of a high ratio of repressor to DNA, all the DNA molecules are covered by repressor molecules and the resulting complexes appear as flattened ribbons with a width of approximately 200 Å. The overall dimensions of these complexes and their substructure indicate that it is very likely that repressor molecules are tightly packed on both “sides” of the DNA helix. Thus two repressor molecules can share the same segment of non-operator DNA by binding to opposite sides of the DNA helix.  相似文献   

15.
The trypsin-resistant core protein of the lac repressor was utilized in protecting operator DNA from two types of enzymatic digestion. Core repressor protects and enhances operator DNA digestion by DNase I in the same fashion as intact repressor, though to a lesser degree on the lower strand. DNase I patterns found for the ternary complexes (protein-sugar-operator) were consistent with the expected affinity alterations of the protein species in response to binding these ligands. The 3′ boundaries obtained by exonuclease III digestion for the intact repressor-operator complex varied slightly from those reported by Shalloway et al. (1980). Asymmetric binding to operator by the core repressor fragment was suggested by differences in the 3′ boundary for the core compared to intact repressor on the promoter-distal side of the complex. A composite picture of repressor structure and function emerges from the protection studies reported here and in the accompanying paper. In light of these and other results, models for repressor binding are examined.  相似文献   

16.
The lac repressor and its tryptic core were studied by ir spectroscopy, and their β-structure content was determined by analysis of the spectra. Using protein-derived reference spectra, we find a β-content for lac repressor of 18% and of 23% for its tryptic core. The higher amount of β-conformation in the tryptic core is confirmed by another type of analysis (decomposition of the spectra in Gaussian curves). These results are discussed with respect to their implications for the structure of the N-terminal “headpiece” of lac repressor and for the mode of interaction of lac repressor with lac operator.  相似文献   

17.
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   

18.
Gene regulatory proteins find their target sites on DNA remarkably quickly; the experimental binding rate for lac repressor is orders-of-magnitude higher than predicted by free diffusion alone. It has been proposed that nonspecific binding aids the search by allowing proteins to slide and hop along DNA. We develop a reaction-diffusion theory of protein translocation that accounts for transport both on and off the strand and incorporates the physical conformation of DNA. For linear DNA modeled as a wormlike chain, the distribution of hops available to a protein exhibits long, power-law tails that make the long-time displacement along the strand superdiffusive. Our analysis predicts effective superdiffusion coefficients for given nonspecific binding and unbinding rate parameters. Translocation rate exhibits a maximum at intermediate values of the binding rate constant, while search efficiency is optimized at larger binding rate constant values. Thus, our theory predicts a region of values of the nonspecific binding and unbinding rate parameters that balance the protein translocation rate and the efficiency of the search. Published data for several proteins falls within this predicted region of parameter values.  相似文献   

19.
Chromatin immunoprecipitation (ChIP) has been used to detect binding of DNA-binding proteins to sites in nuclear and mitochondrial genomes. Here, we describe a method for detecting protein-binding sites on chloroplast DNA, using modifications to the nuclear ChIP procedures. The method was developed using the lac operator (lacO)/lac repressor (LacI) system from Escherichia coli. The lacO sequences were integrated into a single site between the rbcL and accD genes in tobacco plastid DNA and homoplasmic transplastomic plants were crossed with transgenic tobacco plants expressing a nuclear-encoded plastid-targeted GFP-LacI fusion protein. In the progeny, the GFP-LacI fusion protein could be visualized in living tissues using confocal microscopy, and was found to co-localize with plastid nucleoids. Isolated chloroplasts from the lacO/GFP-LacI plants were lysed, treated with micrococcal nuclease to digest the DNA to fragments of ∼600 bp and incubated with antibodies to GFP and protein A-Sepharose. PCR analysis on DNA extracted from the immunoprecipitate demonstrated IPTG (isopropylthiogalactoside)-sensitive binding of GFP-LacI to lacO. Binding of GFP-LacI to endogenous sites in plastid DNA showing sequence similarity to lacO was also detected, but required reversible cross-linking with formaldehyde. This may provide a general method for the detection of binding sites on plastid DNA for specific proteins.  相似文献   

20.
This paper shows that 19F-nuelear magnetic resonance spectroscopy on 3-fluoro-tyrosine and 5-fluorotryptophan-substituted wild-type lactose operon repressors from Escherichia coli can be used to examine the interactions with lac operator DNA.A survey of inducer and salt concentration effects on the repressor-operator complex is presented. The data lead us to a scheme for the interactions between the repressor, operator and inducer, in both binary and ternary complexes, that accommodate the results published by others.The complex between the tetrameric repressor and one 36 base-pair operator DNA fragment results in the simultaneous broadening of the resonances from all four N-terminal DNA binding domains. The actual contacts made by these binding domains are similar but probably not identical.The binding of the inducer molecule to the tetrameric repressor results in an allosteric change that can be monitored by the increased intensity of the resonances from individual tyrosine residues in the N-terminal binding domain. This increased N-terminal tyrosine resonance intensity in the complex is transmitted to repressor subunits that have not yet bound an inducer molecule.  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