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.
Dihydrofolate reductase (5,6,7,8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) was purified from Escherichia coli strains that carried derivatives of the multicopy recombinant plasmid, pJFM8. The results of enzyme kinetic and two-dimensional gel electrophoresis experiments showed that the cloned enzyme is indistinguishable from the chromosomal enzyme. Therefore it can be concluded that these strains are ideal for use as a source of enzyme for further studies on the biochemistry and regulation of this important enzyme. The plasmid derivatives were constructed by recloning experiments that utilized several restriction endonucleases. From the analysis both of these plasmids and the purified dihydrofolate reductase enzymes it was possible to deduce the location and orientation of the dihydrofolate reductase structural gene on the parent plasmid, pJFM8.  相似文献   

2.
3.
Purification and properties of Escherichia coli dihydrofolate reductase.   总被引:5,自引:0,他引:5  
Dihydrofolate reductase has been purified 40-fold to apparent homogeneity from a trimethoprim-resistant strain of Escherichia coli (RT 500) using a procedure that includes methotrexate affinity column chromatography. Determinations of the molecular weight of the enzyme based on its amino acid composition, sedimentation velocity, and sodium dodecyl sulfate gel electrophoresis gave values of 17680, 17470 and 18300, respectively. An aggregated form of the enzyme with a low specific activity can be separated from the monomer by gel filtration; treatment of the aggregate with mercaptoethanol or dithiothreitol results in an increase in enzymic activity and a regeneration of the monomer. Also, multiple molecular forms of the monomer have been detected by polyacrylamide gel electrophoresis. The unresolved enzyme exhibits two pH optima (pH 4.5 and pH 7.0) with dihydrofolate as a substrate. Highest activities are observed in buffers containing large organic cations. In 100 mM imidazolium chloride (pH 7), the specific activity is 47 mumol of dihydrofolate reduced per min per mg at 30 degrees. Folic acid also serves as a substrate with a single pH optimum of pH 4.5. At this pH the Km for folate is 16 muM, and the Vmax is 1/1000 of the rate observed with dihydrofolate as the substrate. Monovalent cations (Na+, K+, Rb+, and Cs+) inhibit dihydrofolate reductase; at a given ionic strength the degree of inhibition is a function of the ionic radius of the cation. Divalent cations are more potent inhibitors; the I50 of BaCl2 is 250 muM, as compared to 125 mM for KCl. Anions neither inhibit nor activate the enzyme.  相似文献   

4.
Folding of dihydrofolate reductase from Escherichia coli   总被引:13,自引:0,他引:13  
The urea-induced equilibrium unfolding transition of dihydrofolate reductase from Escherichia coli was monitored by UV difference, circular dichroism (CD), and fluorescence spectroscopy. Each of these data sets were well described by a two-state unfolding model involving only native and unfolded forms. The free energy of folding in the absence of urea at pH 7.8, 15 degrees C is 6.13 +/- 0.36 kcal mol-1 by difference UV, 5.32 +/- 0.67 kcal mol-1 by CD, and 5.42 +/- 1.04 kcal mol-1 by fluorescence spectroscopy. The midpoints for the difference UV, CD, and fluorescence transitions are 3.12, 3.08, and 3.18 M urea, respectively. The near-coincidence of the unfolding transitions monitored by these three techniques also supports the assignment of a two-state model for the equilibrium results. Kinetic studies of the unfolding and refolding reactions show that the process is complex and therefore that additional species must be present. Unfolding jumps in the absence of potassium chloride revealed two slow phases which account for all of the amplitude predicted by equilibrium experiments. Unfolding in the presence of 400 mM KCl results in the selective loss of the slower phase, implying that there are two native forms present in equilibrium prior to unfolding. Five reactions were observed in refolding: two slow phases designated tau 1 and tau 2 that correspond to the slow phases in unfolding and three faster reactions designated tau 3, tau 4, and tau 5 that were followed by stopped-flow techniques. The kinetics of the recovery of the native form was monitored by following the binding of methotrexate, a tight-binding inhibitor of dihydrofolate reductase, at 380 nm.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Site-specific mutagenesis of dihydrofolate reductase from Escherichia coli   总被引:1,自引:0,他引:1  
Two site-specific mutations of dihydrofolate reductase from Escherichia coli based on the x-ray crystallographic structure were constructed. The first mutation (His-45----Gln) is aimed at assessing the interaction between the imidazole moiety and the pyrophosphate backbone of NADPH. The second (Thr-113----Val) is part of a hydrogen bonding network that contacts the dihydrofolate substrate and may be involved in proton delivery to the N5-C6 imine undergoing reduction. The first mutation was shown to alter both the association and dissociation rate constants for the cofactor so that the dissociation constant was increased 6-40-fold. A corresponding but smaller (fourfold) effect was noted in V/K but not in V compared to the wild-type enzyme. The second was demonstrated to increase the dissociation rate constant for methotrexate 20-30-fold, and presumably dihydrofolate also, with a corresponding 20-30-fold increase in the dissociation constant. In this case an identical effect was noted on V/K but not in V relative to the native enzyme. Thus, in both mutant enzymes the decrease in binding has not been translated into a loss of catalytic efficiency.  相似文献   

6.
The electrostatic potential of Escherichia coli dihydrofolate reductase   总被引:2,自引:0,他引:2  
Escherichia coli dihydrofolate reductase (DHFR) carries a net charge of -10 electrons yet it binds ligands with net charges of -4 (NADPH) and -2 (folate or dihydrofolate). Evaluation and analysis of the electrostatic potential of the enzyme give insight as to how this is accomplished. The results show that the enzyme is covered by an overall negative potential (as expected) except for the ligand binding sites, which are located inside "pockets" of positive potential that enable the enzyme to bind the negatively charged ligands. The electrostatic potential can be related to the asymmetric distribution of charged residues in the enzyme. The asymmetric charge distribution, along with the dielectric boundary that occurs at the solvent-protein interface, is analogous to the situation occurring in superoxide dismutase. Thus DHFR is another case where the shape of the active site focuses electric fields out into solution. The positive electrostatic potential at the entrance of the ligand binding site in E. coli DHFR is shown to be a direct consequence of the presence of three positively charged residues at positions 32, 52, and 57--residues which have also been shown recently to contribute significantly to electronic polarization of the ligand folate. The latter has been postulated to be involved in the catalytic process. A similar structural motif of three positively charged amino acids that gives rise to a positive potential at the entrance to the active site is also found in DHFR from chicken liver, and is suggested to be a common feature in DHFRs from many species. It is noted that, although the net charges of DHFRs from different species vary from +3 to -10, the enzymes are able to bind the same negatively charged ligands, and perform the same catalytic function.  相似文献   

7.
A combination of affinity column chromatography and preparative gel electrophoresis has been used to purify to homogeneity the two isozymes of dihydrofolate reductase from a trimethoprim-resistant strain of Escherichia coli B (RT 500). These enzyme forms are noninterconvertible and are present in crude cell lysates, but other electrophoretic species can be generated durng purification if sulfhydryl-protecting agents, such as dithiothreitol, are not present. The two isozymes, numbered form 1 and form 2 with respect to their decreasing electrophoretic mobilities, have similar molecular weights (18 500), molecular radii (21 A), and apparent Km values for reduced nico inamide adenin- dinucleotide (NADH) and NADH phosphate (NADPH). Both forms contain 2 mol of sulfhydryl/mol of enzyme which can be oxidized to intramolecular disulfide bonds. However, forms 1 and 2 differ physically in their electrophoretic mobility and isoelectric point and kinetically in their pH-activity profile, specific activity, Km for dihydrofolate, and their affinity toward a number of inhibitors.  相似文献   

8.
The interaction of dihydrofolate reductase (EC 1.5.1.3) from Escherichia coli with dihydrofolate and folate analogues has been studied by means of binding and spectroscopic experiments. The aim of the investigation was to determine the number and identity of the binary complexes that can form, as well as pKa values for groups on the ligand and enzyme that are involved with complex formation. The results obtained by ultraviolet difference spectroscopy indicate that, when bound to the enzyme, methotrexate and 2,4-diamino-6,7-dimethylpteridine exist in their protonated forms and exhibit pKa values for their N-1 nitrogens of above 10.0. These values are about five pH units higher than those for the compounds in free solution. The binding data suggest that both folate analogues interact with the enzyme to yield a protonated complex which may be formed by reaction of ionized enzyme with protonated ligand and/or protonated enzyme with unprotonated ligand. The protonated complex formed with 2,4-diamino-6,7-dimethylpteridine can undergo further protonation to form a protonated enzyme-protonated ligand complex, while that formed with methotrexate can ionize to give an unprotonated complex. A group on the enzyme with a pKa value of about 6.3 is involved with the interactions. However, the ionization state of this group has little effect on the binding of dihydrofolate to the enzyme. For the formation of an enzyme-dihydrofolate complex it is essential that the N-3/C-4 amide of the pteridine ring of the substrate be in its neutral form. It appears that dihydrofolate is not protonated in the binary complex.  相似文献   

9.
The internal dynamics of native and immobilized Escherichia coli dihydrofolate reductase (DHFR) have been examined using incoherent quasielastic neutron scattering. These results reveal no difference between the high frequency vibration mean-square displacement of the native and the immobilized E. coli DHFR. However, length-scale-dependent, picosecond dynamical changes are found. On longer length scales, the dynamics are comparable for both DHFR samples. On shorter length scales, the dynamics is dominated by local jump motions over potential barriers. The residence time for the protons to stay in a potential well is tau = 7.95 +/- 1.02 ps for the native DHFR and tau = 20.36 +/- 1.80 ps for the immobilized DHFR. The average height of the potential barrier to the local motions is increased in the immobilized DHFR, and may increase the activation energy for the activity reaction, decreasing the rate as observed experimentally. These results suggest that the local motions on the picosecond timescale may act as a lubricant for those associated with DHFR activity occurring on a slower millisecond timescale. Experiments indicate a significantly slower catalytic reaction rate for the immobilized E. coli DHFR. However, the immobilization of the DHFR is on the exterior of the enzyme and essentially distal to the active site, thus this phenomenon has broad implications for the action of drugs distal to the active site.  相似文献   

10.
11.
A kinetic mechanism is presented for Escherichia coli dihydrofolate reductase which describes the full time course of the enzymatic reaction over a wide range of substrate and enzyme concentrations at pH 7.2 and 20 degrees C. Specific rate constants were estimated by computer simulation of the full time course of single turnover, burst, and steady-state experiments using both nondeuterated and deuterated NADPH. The mechanism involves the random addition of substrates, but the substrates and enzyme are not at equilibrium prior to the chemical transformation step. The rate-limiting step follows the chemical transformation, and the maximum velocity of the reaction is limited by the release of the product tetrahydrofolate. The full time course of the reaction is markedly affected by the formation of the enzyme-NADPH-tetrahydrofolate abortive complex, but not by the enzyme-NADP-dihydrofolate abortive complex.  相似文献   

12.
As a continuation to our studies on the importance of interloop interactions in the Escherichia coli DHFR catalytic cycle, we have investigated the role of the betaG-betaH loop in modulating the closed and occluded conformations of the Met20 loop during the DHFR catalytic cycle. Specifically, to assess the importance of the hydrogen bond formed between Ser148 in the betaG-betaH loop and the Met20 loop, Ser148 was independently substituted with aspartic acid, alanine, and lysine. Moreover, the betaG-betaH loop was deleted entirely to yield the Delta(146-148) DHFR mutant. Steady-state turnover rates for all mutants were at most 3-fold lower than the wild-type rate. Lack of an isotope effect on this rate indicated the chemistry step does not contribute to the steady-state turnover. Consistent with this finding, hydride transfer rates for the DHFR mutants were at least 10-fold greater than the observed steady-state rates. The values ranged from a 30% decrease (Ser148Ala and Ser148Lys) to a 50% increase (Ser148Asp) in rate relative to that of the wild type. Modifications of the betaG-betaH loop enhanced the affinity for the cofactor and decreased the affinity for pterin, as determined by the K(D) values of the mutant proteins. Further analysis of Ser148Ala and Delta(146-148) DHFRs indicated these effects were manifest mainly in ligand off rates, although in some cases the on rate was affected. The Ser148Asp and Delta(146-148) mutations perturbed the preferred catalytic cycle through the introduction of branching at key intermediates. Rather than following the single WT pathway which involves loss of NADP(+) and rebinding of NADPH to precede loss of the product H4F (negative cooperativity), the mutants can reenter the catalytic cycle through different pathways. These findings suggest that the role of the interloop interaction between the betaG-betaH loop and the Met20 loop is to modulate ligand off rates allowing for proper cycling through the preferred kinetic pathway.  相似文献   

13.
14.
The Escherichia coli strain carrying pTP 6-10 which was constructed in our previous work (Iwakura, M., et al. (1983) J. Biochem. 93, 927-930) produces more than 400-fold dihydrofolate reductase as compared with the strain without the plasmid. Dihydrofolate reductase was highly purified from the cell-free extract of the plasmid strain simply by two steps; ammonium sulfate fractionation and ion-exchange chromatography. By 10-fold purification, the enzyme was essentially homogeneous as judged by polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate. The restriction map of pTP 6-10 was also determined and the plasmid was shown to have an Ava I, an EcoR I, a Pst I, a Pvu I, and a Pvu II site. Our results indicate that the plasmid strain is suitable as a source of the enzyme and that plasmid pTP 6-10 is promising as a versatile plasmid vector for efficiently yielding the product of the cloned gene.  相似文献   

15.
Full time course studies of the kinetic activity of Escherichia coli dihydrofolate reductase show that there is an increase in activity with time. The half-time for this hysteretic behavior is about 9 s. Preincubation of the enzyme with either of the substrates abolishes the lag and results in initial velocities which are 2-2.3-fold faster than those observed for the non-preincubated enzyme. The kinetic properties of the activated and nonactivated forms of the enzyme appear to be similar as measured by the full time course of the reaction. The results are consistent with observations for NADPH binding studies that the enzyme exists in two interconvertible forms, one of which is incapable of binding NADPH (Cayley, P. J., Dunn, S. M. J., and King, R. W. (1981) Biochemistry 20, 874-879).  相似文献   

16.
Summary Culture of the thymine-dependent mutant of Rhizobium trifolii T37 was synchronized with phenethanol. During bacterial synchronous growth with synchronized replication of DNA, cells were differentially labeled using subsequently 3H-thymidine of low and high specific activity. The grain tracks produced in autoradiographs of chromosomes were denser on both ends than in the middle. In control experiments with bacteria labeled only with 3H-thymidine of low specific activity, the grain density was uniform throughout the grain track. The results constitute clear evidence of bidirectional replication of R. trifolli chromosome.  相似文献   

17.
To elucidate the effects of pressure on the function of Escherichia coli dihydrofolate reductase (DHFR), the enzyme activity and the dissociation constants of substrates and cofactors were measured at pressures up to 250 MPa at 25 degrees C and pH 7.0. The enzyme activity decreased with increasing pressure, accompanying the activation volume of 7.8 ml mol(-1). The values of the Michaelis constant (K(m)) for dihydrofolate and NADPH were slightly higher at 200 MPa than at atmospheric pressure. The hydride-transfer step was insensitive to pressure, as monitored by the effects of the deuterium isotope of NADPH on the reaction velocity. The dissociation constants of substrates and cofactors increased with pressure, producing volume reductions from 6.5 ml mol(-1) (tetrahydrofolate) to 33.5 ml mol(-1) (NADPH). However, the changes in Gibbs free energy with dissociation of many ligands showed different pressure dependences below and above 50 MPa, suggesting conformational changes of the enzyme at high pressure. The enzyme function at high pressure is discussed based on the volume levels of the intermediates and the candidates for the rate-limiting process.  相似文献   

18.
An expression vector, pKGR, for the gor gene from Escherichia coli encoding glutathione reductase was constructed by subcloning of an AvaII fragment of the Clarke & Carbon bank plasmid pGR [Greer & Perham (1986) Biochemistry 25, 2736-2742] into the plasmid pKK223-3. The expression of glutathione reductase from the plasmid pKGR was found to have been successfully placed under the control of the tac promoter. Transformation of E. coli cells with this plasmid resulted in 100-200-fold increase in glutathione reductase activity in cell-free extracts. A rapid purification procedure for the enzyme, based on affinity chromatography on Procion Red HE-7B-CL-Sepharose 4B, was developed. The purified enzyme was homogeneous as judged by SDS/polyacrylamide-gel electrophoresis, and all its properties were consistent with the DNA sequence of the gene [Greer & Perham (1986) Biochemistry 25, 2736-2742] and with those previously reported for E. coli glutathione reductase [Mata, Pinto & Lopez-Barea (1984) Z. Naturforsch. C. Biosci. 39, 908-915]. These experiments have enabled an investigation of the protein chemical and mechanistic properties of the enzyme by site-directed mutagenesis.  相似文献   

19.
Escherichia coli (thyA DeltafolA) mutants are viable and can grow in minimal medium when supplemented with thymidine alone. Here we present evidence from in vivo and in vitro studies that the ydgB gene determines an alternative dihydrofolate reductase that is related to the trypanosomatid pteridine reductases. We propose to rename this gene folM.  相似文献   

20.
Product identification and kinetic data are presented for the conversion of 7,8-dihydrobiopterin into tetrahydrobiopterin by purified rabbit brain dihydrofolate reductase.  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