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 203 毫秒
1.
The conformation changes of dihydrofolate reductase (DHFR) from chicken liver in guanidine hy-drochloride were monitored by protein intrinsic fluorescence, hydrophobic fluorescence probe TNS and limited proteol-ysis by proteinase K. The kinetics of the enzyme denaturation were also studied and compared with its activity changes. It was indicated by the enhanced fluorescence of 2-p-toluidinylnaphthalene (TNS) that a subtle conforma-tional change of the enzyme in dilute GuHCl parallels GuHCl-induced activation. At GuHCl concentration higher than 0.75 mol/L, the conformational change can be detected by increased susceptibility of the enzyme to proteinase K, but no significant gross conformational change of the enzyme molecule is observed by intrinsic fluorescence up to a GuHCl concentration of 1.2 mol/L. The results suggest that the denaturation of DHFR by GuHCl does not follow strictly the two-state model. The enzyme seems to open up sequentially with increasing concentrations of denaturants, mainly at th  相似文献   

2.
Weikl TR  Boehr DD 《Proteins》2012,80(10):2369-2383
Protein function often involves changes between different conformations. Central questions are how these conformational changes are coupled to the binding or catalytic processes during which they occur, and how they affect the catalytic rates of enzymes. An important model system is the enzyme dihydrofolate reductase (DHFR) from Escherichia coli, which exhibits characteristic conformational changes of the active‐site loop during the catalytic step and during unbinding of the product. In this article, we present a general kinetic framework that can be used (1) to identify the ordering of events in the coupling of conformational changes, binding, and catalysis and (2) to determine the rates of the substeps of coupled processes from a combined analysis of nuclear magnetic resonance R2 relaxation dispersion experiments and traditional enzyme kinetics measurements. We apply this framework to E. coli DHFR and find that the conformational change during product unbinding follows a conformational‐selection mechanism, that is, the conformational change occurs predominantly prior to unbinding. The conformational change during the catalytic step, in contrast, is an induced change, that is, the change occurs after the chemical reaction. We propose that the reason for these conformational changes, which are absent in human and other vertebrate DHFRs, is robustness of the catalytic rate against large pH variations and changes to substrate/product concentrations in E. coli. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
Ensemble kinetics and single-molecule fluorescence microscopy were used to study conformational transitions associated with enzyme catalysis by dihydrofolate reductase (DHFR). The active site loop of DHFR was labeled with a fluorescence quencher, QSY35, at amino acid position 17, and the fluorescent probe, Alexa555, at amino acid 37, by introducing cysteines at these sites with site-specific mutagenesis. The distance between the probes was such that approximately 50% fluorescence resonance energy transfer (FRET) occurred. The double-labeled enzyme retained essentially full catalytic activity, and stopped-flow studies of both the forward and reverse reactions revealed that the distance between probes increased prior to hydride transfer. A fluctuation in fluorescence intensity of single molecules of DHFR was observed in an equilibrium mixture of substrates but not in their absence. Ensemble rate constants were derived from the distributions of lifetimes observed and attributed to a reversible conformational change. Studies were carried out with both NADPH and NADPD as substrates, with no measurable isotope effect. Similar studies with a G121V mutant DHFR resulted in smaller rate constants. This mutant DHFR has reduced catalytic activity, so that the collective data for the conformational change suggest that the conformational change being observed is associated with catalysis and probably represents a conformational change prior to hydride transfer. If the change in fluorescence is attributed to a change in FRET, the distance change associated with the conformational change is approximately 1-2 A. These results are correlated with other measurements related to conformation coupled catalysis.  相似文献   

4.
R67 dihydrofolate reductase (DHFR) shares no sequence or structural homology with chromosomal DHFRs. This enzyme arose recently in response to the clinical use of the antibacterial drug trimethoprim. R67 DHFR is a homotetramer possessing a single active site pore. A high-resolution crystal structure shows the homotetramer possesses exact 222 symmetry [Narayana, N., et al. (1995) Nat. Struct. Biol. 2, 1018-1025]. This symmetry dictates four symmetry-related binding sites must exist for each substrate as well as each cofactor. Isothermal titration calorimetry studies, however, indicate only two molecules bind: either two dihydrofolate molecules, two NADPH molecules, or one substrate and one cofactor [Bradrick, T. D., et al. (1996) Biochemistry 35, 11414-11424]. The latter is the productive ternary complex. To evaluate the role of S65, Q67, I68, and Y69 residues, located near the center of the active site pore, site-directed mutagenesis was performed. One mutation in the gene creates four mutations per active site pore which typically result in large cumulative effects. Steady state kinetic data indicate the mutants have altered K(m) values for both cofactor and substrate. For example, the Y69F R67 DHFR displays an 8-fold increase in the K(m) for dihydrofolate and a 20-fold increase in the K(m) for NADPH. Residues involved in ligand binding in R67 DHFR display very little, if any, specificity, consistent with their possessing dual roles in binding. These results support a model where R67 DHFR utilizes an unusual "hot spot" binding surface capable of binding both ligands and indicate this enzyme has adopted a novel yet simple approach to catalysis.  相似文献   

5.
Hicks SN  Smiley RD  Hamilton JB  Howell EE 《Biochemistry》2003,42(36):10569-10578
R67 dihydrofolate reductase (DHFR), which catalyzes the NADPH dependent reduction of dihydrofolate to tetrahydrofolate, belongs to a type II family of R-plasmid encoded DHFRs that confer resistance to the antibacterial drug trimethoprim. Crystal structure data reveals this enzyme is a homotetramer that possesses a single active site pore. Only two charged residues in each monomer are located near the pore, K32 and K33. Site-directed mutants were constructed to probe the role of these residues in ligand binding and/or catalysis. As a result of the 222 symmetry of this enzyme, mutagenesis of one residue results in modification at four related sites. All mutants at K32 affected the quaternary structure, producing an inactive dimer. The K33M mutant shows only a 2-4-fold effect on K(m) values. Salt effects on ligand binding and catalysis for K33M and wildtype R67 DHFRs were investigated to determine if these lysines are involved in forming ionic interactions with the negatively charged substrates, dihydrofolate (overall charge of -2) and NADPH (overall charge of -3). Binding studies indicate that two ionic interactions occur between NADPH and R67 DHFR. In contrast, the binding of folate, a poor substrate, to R67 DHFR.NADPH appears weak as a titration in enthalpy is lost at low ionic strength. Steady-state kinetic studies for both wild type (wt) and K33M R67 DHFRs also support a strong electrostatic interaction between NADPH and the enzyme. Interestingly, quantitation of the observed salt effects by measuring the slopes of the log of ionic strength versus the log of k(cat)/K(m) plots indicates that only one ionic interaction is involved in forming the transition state. These data support a model where two ionic interactions are formed between NADPH and symmetry related K32 residues in the ground state. To reach the transition state, an ionic interaction between K32 and the pyrophosphate bridge is broken. This unusual scenario likely arises from the constraints imposed by the 222 symmetry of the enzyme.  相似文献   

6.
Despite much work, many key aspects of the mechanism of the dihydrofolate reductase (DHFR) catalyzed reduction of dihydrofolate remain unresolved. In bacterial forms of DHFR both substrate and water access to the active site are controlled by the conformation of the mobile M20 loop. In vertebrate DHFRs only one conformation of the residues corresponding to the M20 loop has been observed. Access to the active site was proposed to be controlled by residue 31. MD simulations of chicken DHFR complexed with substrates and cofactor revealed a closing of the side chain of Tyr 31 over the active site on binding of dihydrofolate. This conformational change was dependent on the presence of glutamate on the para-aminobenzoylamide moiety of dihydrofolate. In its absence, the conformation remained open. Although water could enter the active site and hydrogen bond to N5 of dihydrofolate, indicating the feasibility of water as the proton donor, this was not controlled by the conformation of Tyr 31. The water accessibility of the active site was low for both conformations of Tyr 31. However, when hydride was transferred from NADPH to C6 of dihydrofolate before protonation, the average time during which water was found in hydrogen bonding distance to N5 of dihydrofolate in the active site increased almost fivefold. These results indicated that water can serve as the Broensted acid for the protonation of N5 of dihydrofolate during the DHFR catalyzed reduction.  相似文献   

7.
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   

8.
Four spin-labeled inhibitors of dihydrofolate reductase (DHFR) have been synthesized, each of which has the 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) reporting group at a different distance from the 2,4-diaminopyrimidine moiety by which the inhibitors are anchored and oriented in the active site. Inhibitors in which the TEMPO group is attached by a short side chain are weakly bound to DHFR from bacteria (Streptococcus faecium and Lactobacillus casei), to the bovine enzyme and to recombinant human DHFR. However, binding is sufficiently tight, especially in the ternary complexes with NADPH, for recording of the EPR spectra of the bound ligands. The spectra indicate that when these inhibitors are bound to the enzyme the TEMPO group is highly immobilized with correlation time, tau c, 4-20ns. Inhibitors that have the reporter group attached to the glutamate moiety of methotrexate bind to all four DHFRs more tightly than the inhibitors with shorter side chains by factors of up to 10(6). However, in most complexes formed by the inhibitors with longer side chains immobilization of the TEMPO group is slight (tau c 0.2-4 ns). These results are in general agreement with predictions from X-ray crystallographic results including thermal factors but there are some unanticipated differences between some results for bacterial and eukaryotic enzymes. Three of the splin-labeled inhibitors would provide good probes for distance measurements in and around the active site of mammalian DHFR.  相似文献   

9.
Plasmid-encoded bacterial R67 dihydrofolate reductase (DHFR) is a NADPH-dependent enzyme unrelated to chromosomal DHFR in amino acid sequence and structure. R67 DHFR is insensitive to the bacterial drug trimethoprim in contrast to chromosomal DHFR. The crystal structure of Q67H mutant of R67 DHFR bound to NADP(+) has been determined at 1.15 angstroms resolution. The cofactor assumes an extended conformation with the nicotinamide ring bound near the center of the active site pore, the ribose and pyrophosphate group (PP(i)) extending toward the outer pore. The ribonicotinamide exhibits anti conformation as in chromosomal DHFR complexes. The relative orientation between the PP(i) and the nicotinamide ribose differs from that observed in chromosomal DHFR-NADP(+) complexes. The coenzyme displays symmetrical binding mode with several water-mediated hydrogen bonds with the protein besides ionic, stacking, and van der Waals interactions. The structure provides a molecular basis for the observed stoichiometry and cooperativity in ligand binding. The ternary model based on the present structure and the previous R67 DHFR-folate complex provides insight into the catalytic mechanism and indicates that the relative orientation of the reactants in plasmid DHFR is different from that seen in chromosomal DHFRs.  相似文献   

10.
Structural differences in dihydrofolate reductases from different species have been exploited to develop specific inhibitory molecules, such as chemotherapeutic agents, antibiotics or antihelminthics, that show species specificity or selectivity. As dihydrofolate reductase (DHFR) is a crucial enzyme for the synthesis of purines, pyrimidines and some amino acids, and also because developing insects show a remarkably rapid rate of cell division, DHFR is a potentially promising target for the discovery of novel insecticides. We have thus isolated and characterized the enzyme from a serious agricultural pest, Heliothis (Helicoverpa) virescens, the tobacco budworm. Sequencing tryptic peptides of the 35 000-fold purified DHFR allowed the subsequent isolation of a partial cDNA, with the full Dhfr gene sequence obtained from a genomic library. The H. virescens Dhfr spans 4 kb, with three introns, and encodes 185 amino acids. The enzyme shows an overall similarity of approximately 68% with DHFR from other metazoans, which has facilitated the molecular modeling of the protein. DHFRs from insects appear to have strikingly reduced sensitivity to inhibition by methotrexate, compared with the vertebrate enzymes, and this reduction was also reflected in the total binding energy seen after modeling experiments. Four residues that may be characteristic of insect DHFR, as well as a unique cysteine in the H. virescens DHFR active site, offer insight into the nature of inhibitor selectivity and provide suitable target sites for insecticide discovery.  相似文献   

11.
Rapid, polyamine-induced degradation of mammalian ornithine decarboxylase (L-ornithine carboxy-lyase, EC 4.1.1.17) (ODC) is though to be controlled by the availability of a small, ODC-binding protein termed antizyme. In this study we have investigated the ability of antizyme to bind ODC protein in various altered physiological states. In particular, cold, NaCl, spermidine and deprivation of coenzyme and substrate enhance enzyme-antizyme complex formation and are all found to promote ODC homodimer dissociation. Conversely, conditions that maintain the active ODC homodimer state prevent antizyme binding and inactivation of ODC. Further, covalent modification of ODC near its active site by difluoromethylornithine or phosphate also increases its sensitivity to antizyme. These results suggest that the initial signal in ODC degradation may actually be a subtle conformational change in the enzyme that enables antizyme to bind to the enzyme and may subsequently facilitate its degradation.  相似文献   

12.
Escherichia coli RNA polymerase loses 55-65% of its catalytic activity on reaction with Nbf-Cl (4-choro-7-nitrobenzofurazan). This partial inactivation was shown to be the result of specific impairment of RNA-chain elongation, since initiation of RNA chains was not altered after treatment with Nbf-Cl. The site of reaction was shown to be a unique thiol on the beta-subunit. This thiol is not accessible to reaction with 5,5'-dithiobis-(2-nitrobenzoic acid). No protection of the enzyme against reaction with Nbf-Cl could be obtained with the inhibitor rifamycin nor with calf thymus DNA, GTP or 1,10-phenanthroline, indicating that the unique thiol is probably not within the active site. The specific impairment of RNA-chain elongation thus appears to be the result of a local conformational change which leaves chain initiation unimpaired. Changes observed in the tryptophan fluorescence spectrum of the enzyme or reaction with Nbf-Cl are consistent with formation of a Meisenheimer complex of the reagent with a nucleophilic group on the enzyme near the reactive thiol. It is proposed that formation of such a complex and a subsequent conformational change renders this thiol unusually susceptible to reaction with Nbf-Cl.  相似文献   

13.
The kinetic theory of the substrate reaction during modification of enzyme activity previously described by Tsou [Tsou (1988),Adv. Enzymol. Relat. Areas Mol. Biol. 61, 381–436] has been applied to a study of the kinetics of the course of inactivation of the mitochondrial succinate-ubiquinone reductase by 5,5′-dithiobis-(2-nitro-benzoic acid) (DTNB). The results show that the inactivation of this enzyme by DTNB is a conformation-change-type inhibition which involves a conformational change of the enzyme before inactivation. The microscopic rate constants were determined for the reaction of the inactivator with the enzyme. The presence of the substrate provides marked protection of this enzyme against inactivation by DTNB. The modification reaction of the enzyme using DTNB was shown to follow a triphasic course by following the absorption at 412 nm. Among these reactive thiol groups, the fast-reaction thiol group is essential for the enzyme activity. The results suggest that the essential thiol group is situated at the succinate-binding site of the mitochondrial succinate-ubiquinone reductase.  相似文献   

14.
The kinetic characteristics of a purified insect dihydrofolate reductase (DHFR) have been described. The Km values for the substrate dihydrofolate and the cofactor NADPH have been estimated by primary and secondary Hanes plots to be 0.3 and 5.2 microM, respectively. Drosophila melanogaster DHFR can use folate and NADH at acidic pH values, but at a much lower rate than the preferred substrate and cofactor. Folic acid is a partial competitive inhibitor of Drosophila DHFR (Ki = 0.4 microM) and trimethoprim is a complete competitive inhibitor (Ki = 5.4 microM). Methotrexate binds less tightly to the Drosophila enzyme than to many other DHFRs (Kd = 0.9 nM). Drosophila DHFR is inhibited by KCl and organic mercurials and is slightly activated by urea. These data indicate that Drosophila DHFR has some characteristics which are typical of vertebrate DHFRs and others which are typical of prokaryotic DHFRs. The study of this enzyme, therefore, should aid in the definition of the structural features that are responsible for the kinetic characteristics in different DHFRs.  相似文献   

15.
Dihydrofolate reductase as a therapeutic target   总被引:9,自引:0,他引:9  
The folate antagonists are an important class of therapeutic compounds, as evidenced by their use as antiinfective, antineoplastic, and antiinflammatory drugs. Thus far, all of the clinically useful drugs of this class have been inhibitors of dihydrofolate reductase (DHFR), a key enzyme in the synthesis of thymidylate, and therefore, of DNA. The basis of the antiinfective selectivity of these compounds is clear; the antifolates trimethoprim and pyrimethamine are potent inhibitors of bacterial and protozoal DHFRs, respectively, but are only weak inhibitors of mammalian DHFRs. These species-selective agents apparently exploit the differences in the active site regions of the parasite and host enzymes. Methotrexate is the DHFR inhibitor used most often in a clinical setting as an anticancer drug and as an antiinflammatory and immunosuppressive agent. Considerable progress has been made recently in understanding the biochemical basis for the selectivity of this drug and the biochemical mechanism (or mechanisms) responsible for the development of resistance to treatment with the drug. This understanding has led to a new generation of DHFR inhibitors that are now in clinical trials.  相似文献   

16.
All of the 181 possible overlapping hexapeptides as well as 179 octapeptides covering the amino acid sequence of human dihydrofolate reductase (hDHFR) were synthesized on polyethylene supports. The synthetic procedure of Geysen et al. (Geysen, H. M., Rodda, S. J., Mason, T. J., Tribbick, G., and Schoofs, P. G. (1987) J. Immunol. Methods 102, 259-274) was modified to obtain up to 100 nmol of peptide on each pin. Peptides constituting antigenic epitopes on hDHFR were identified by examining the binding of antibodies raised against both native and denatured hDHFR to these peptides by enzyme-linked immunosorbent assay. The peptides bound in a similar pattern to polyclonal antibodies against both native and denatured dihydrofolate reductase (DHFR). Six major epitopes were located corresponding to residues 27-33, 45-51, 67-74, 133-139, 153-158, and 176-181 using both hexapeptides and octapeptides. An additional epitope, constituting residues 14-21, was found by the use of octapeptides. Most of the epitopes are hydrophilic and reside largely in "loop" regions at the boundaries of secondary structural elements of hDHFR. This observation is consistent with our previous results which suggested that ligand binding at the active site of the enzyme can cause a dramatic reduction in antibody binding to DHFR due to conformational constraints in flexible loop regions in various parts of the molecule. The similarity of the immunogenic profiles of native versus denatured hDHFR indicates that the two forms of the antigen share the same amino acid sequence-specific epitopes. Competitive enzyme-linked immunosorbent assay showed that the binding of anti-hDHFR antiserum to both native and denatured hDHFR was inhibited by approximately 30% by the seven antigenic peptides, indicating that a significant proportion of the antibodies elicited by this enzyme is specific for short peptides. Besides revealing the antigenic structure of DHFR our results provide a rational basis for the design of mutant DHFRs to study the importance of loop residues in the conformational dynamics of the enzyme.  相似文献   

17.
Polyclonal antibodies against dihydrofolate reductase (DHFR) from the human lymphoblastoid cell line WIL-2/M4 were used as probes to compare the antigenic structures in solution of native DHFRs obtained from a broad range of species and their complexes with substrate, cofactor, and folate antagonist inhibitors. All these antibodies could bind to the denatured human DHFR, indicating that they were specific for the primary structure of this enzyme. Denatured chicken liver and L1210 murine leukemic DHFRs competed for all of the antibodies that bound to the human enzyme, although less effectively than the denatured human enzyme, showing the presence of similar epitopes among the vertebrate enzymes. However, both direct binding and competition experiments showed low antibody cross-reactivities with native chicken liver (8%) and murine (10%) DHFRs, suggesting differences in the disposition of similar epitopes in these enzymes. The lactobacillus casei DHFR showed a low amount (less than 2%) of cross-reactivity with the antibodies while the same antibodies did not cross-react with the Escherichia coli enzyme. DHFR from soybean seedlings competed for a large proportion (70%) of the anti-human DHFR antibodies, indicating a close similarity in the antigenic structures of plant and animal DHFRs. Binary complexes of the L. casei, avian, murine, and human DHFRs with dihydrofolate, methotrexate (MTX), trimethoprim (TMP), NADPH, and NADP+ all showed significantly lower antibody binding capacity as compared with the corresponding free enzymes. Further, these ligands inhibited antibody binding to the enzyme to varying degrees.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
To examine whether dihydrofolate reductase (DHFR) from deep-sea bacteria has undergone molecular evolution to adapt to high-pressure environments, we cloned eight DHFRs from Shewanella species living in deep-sea and ambient atmospheric-pressure environments, and subsequently purified six proteins to compare their structures, stabilities, and functions. The DHFRs showed 74–90% identity in primary structure to DHFR from S. violacea, but only 55% identity to DHFR from Escherichia coli (ecDHFR). Far-ultraviolet circular dichroism and fluorescence spectra suggested that the secondary and tertiary structures of these DHFRs were similar. In addition, no significant differences were found in structural stability as monitored by urea-induced unfolding and the kinetic parameters, K m and k cat; although the DHFRs from Shewanella species were less stable and more active (2- to 4-fold increases in k cat/K m) than ecDHFR. Interestingly, the pressure effects on enzyme activity revealed that DHFRs from ambient-atmospheric species are not necessarily incompatible with high pressure, and DHFRs from deep-sea species are not necessarily tolerant of high pressure. These results suggest that the DHFR molecule itself has not evolved to adapt to high-pressure environments, but rather, those Shewanella species with enzymes capable of retaining functional activity under high pressure migrated into the deep-sea.  相似文献   

19.
R67 dihydrofolate reductase (DHFR) is a type II DHFR produced by bacteria as a resistance mechanism to the increased clinical use of the antibacterial drug trimethoprim. Type II DHFRs are not homologous in either sequence or structure with chromosomal DHFRs. The type II enzymes contain four identical subunits which form a homotetramer containing a single active site pore accessible from either end. Although the crystal structure of the complex of R67 DHFR with folate has been reported [Narayana et al. (1995) Nat. Struct. Biol. 2, 1018], the nature of the ternary complex which must form with substrate and cofactor is unclear. We have performed transferred NOE and interligand NOE (ILOE) studies to analyze the ternary complexes formed from NADP(+) and folate in order to probe the structure of the ternary complex. Consistent with previous studies of the binary complex formed from another type II DHFR, the ribonicotinamide bond of NADP(+) was found to adopt a syn conformation, while the adenosine moiety adopts an anti conformation. Large ILOE peaks connecting NADP(+) H4 and H5 with folate H9 protons are observed, while the absence of a large ILOE connecting NADP(+) H4 and H5 with folate H7 indicates that the relative orientation of the two ligands differs significantly from the orientation in the chromosomal enzyme. To obtain more detailed insight, we prepared and studied the folate analogue 2-deamino-2-methyl-5,8-dideazafolate (DMDDF) which contains additional protons in order to provide additional NOEs. For this analogue, the exchange characteristics of the corresponding ternary complex were considerably poorer, and it was necessary to utilize higher enzyme concentrations and higher temperature in order to obtain ILOE information. The results support a structure in which the NADP(+) and folate/DMDDF molecules extend in opposite directions parallel to the long axis of the pore, with the nicotinamide and pterin ring systems approximately stacked at the center. Such a structure leads to a ternary complex which is in many respects similar to the gas-phase theoretical calculations of the dihydrofolate-NADPH transition state by Andres et al. [(1996) Bioorg. Chem. 24, 10-18]. Analogous NMR studies performed on folate, DMDDF, and R67 DHFR indicate formation of a ternary complex in which two symmetry-related binding sites are occupied by folate and DMDDF.  相似文献   

20.
    
The kinetic theory of the substrate reaction during modification of enzyme activity previously described by Tsou [Tsou (1988),Adv. Enzymol. Relat. Areas Mol. Biol. 61, 381–436] has been applied to a study of the kinetics of the course of inactivation of the mitochondrial succinate-ubiquinone reductase by 5,5-dithiobis-(2-nitro-benzoic acid) (DTNB). The results show that the inactivation of this enzyme by DTNB is a conformation-change-type inhibition which involves a conformational change of the enzyme before inactivation. The microscopic rate constants were determined for the reaction of the inactivator with the enzyme. The presence of the substrate provides marked protection of this enzyme against inactivation by DTNB. The modification reaction of the enzyme using DTNB was shown to follow a triphasic course by following the absorption at 412 nm. Among these reactive thiol groups, the fast-reaction thiol group is essential for the enzyme activity. The results suggest that the essential thiol group is situated at the succinate-binding site of the mitochondrial succinate-ubiquinone 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