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.
Dihydrofolate reductase (DHFR) from extracts of Mycobacterium smegmatis strain mc2(6) and trimethoprim-resistant mutant mc2(26) was purified to homogeneity. In crude extracts, the specific activity of the enzyme from the trimethoprim resistant strain was comparable to that from the sensitive strain. The DHFR from both sources was purified using affinity chromatography on MTX-Sepharose followed by Mono Q FPLC. The enzyme has an apparent molecular mass of 23 kDa from gel filtration on Sephadex G-100 and from SDS-PAGE. Amino terminal sequence analysis showed homology with DHFRs from a subset of other gram-positive organisms. The purified enzyme from the trimethoprim-sensitive organism exhibited Km values for H2folate and NADPH of 0.68 +/- 0.2 microM and 21 +/- 4 microM, respectively. The Km values for H2folate and NADPH for the enzyme from the drug-resistant organism were 1.8 +/- 0.4 microM and 5.3 +/- 1.5 microM, respectively. A kcat of 4.5 sec-1 was determined for the DHFR from both sources. The enzyme from both sources was competitively inhibited by pyrimethamine and trimethoprim. The Ki value of trimethoprim, for the enzyme from the drug-resistant organism was about six-fold higher than for the enzyme from drug-sensitive strain. Our data suggest that mutation of DHFR contributes to trimethoprim resistance in the mc2(26) strain of M. smegmatis.  相似文献   

2.
Mass spectrometry technologies for measurement of cellular metabolism are opening new avenues to explore drug activity. Trimethoprim is an antibiotic that inhibits bacterial dihydrofolate reductase (DHFR). Kinetic flux profiling with (15)N-labeled ammonia in Escherichia coli reveals that trimethoprim leads to blockade not only of DHFR but also of another critical enzyme of folate metabolism: folylpoly-gamma-glutamate synthetase (FP-gamma-GS). Inhibition of FP-gamma-GS is not directly due to trimethoprim. Instead, it arises from accumulation of DHFR's substrate dihydrofolate, which we show is a potent FP-gamma-GS inhibitor. Thus, owing to the inherent connectivity of the metabolic network, falling DHFR activity leads to falling FP-gamma-GS activity in a domino-like cascade. This cascade results in complex folate dynamics, and its incorporation in a computational model of folate metabolism recapitulates the dynamics observed experimentally. These results highlight the potential for quantitative analysis of cellular metabolism to reveal mechanisms of drug action.  相似文献   

3.
The Escherichia coli dihydrofolate reductase (DHFR) gene has been used as a genetic marker specifying trimethoprim resistance (TmpR). In order to use the DHFR gene as a versatile expression marker, we have constructed three types of plasmids: promoter cloning vector, terminator cloning vector, and the plasmid containing the DHFR gene cassette. In these systems, the selection of recombinant plasmids was carried out just by examining the TmpR phenotype of the transformed cells. Then, levels of the enzymatic activity of DHFR were measured to evaluate the efficiency of promoters and terminators in the fused DNA fragment. An expression plasmid which resulted in the E. coli host cells being able to produce DHFR up to 20% of total cellular proteins was also constructed by changing the promoter and Shine-Dalgarno sequences of the DHFR gene.  相似文献   

4.
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   

5.
The relationship between sequence variation and phenotype is poorly understood. Here, we use metabolomic analysis to elucidate the molecular mechanism underlying the filamentous phenotype of E. coli strains that carry destabilizing mutations in dihydrofolate reductase (DHFR). We find that partial loss of DHFR activity causes reversible filamentation despite SOS response indicative of DNA damage, in contrast to thymineless death (TLD) achieved by complete inhibition of DHFR activity by high concentrations of antibiotic trimethoprim. This phenotype is triggered by a disproportionate drop in intracellular dTTP, which could not be explained by drop in dTMP based on the Michaelis–Menten‐like in vitro activity curve of thymidylate kinase (Tmk), a downstream enzyme that phosphorylates dTMP to dTDP. Instead, we show that a highly cooperative (Hill coefficient 2.5) in vivo activity of Tmk is the cause of suboptimal dTTP levels. dTMP supplementation rescues filamentation and restores in vivo Tmk kinetics to Michaelis–Menten. Overall, this study highlights the important role of cellular environment in sculpting enzymatic kinetics with system‐level implications for bacterial phenotype.  相似文献   

6.
Dihydrofolate reductase (DHFR) catalyzes the NADPH-dependent reduction of dihydrofolate to tetrahydrofolate and is essential for the synthesis of thymidylate, purines and several amino acids. Inhibition of the enzyme's activity leads to arrest of DNA synthesis and cell death. The enzyme has been studied extensively as a drug target for bacterial, protozoal and fungal infections, and also for neoplastic and autoimmune diseases. Here, we report the crystal structure of dihydrofolate reductase from Mycobacterium tuberculosis, a human pathogen responsible for the death of millions of human beings per year. Three crystal structures of ternary complexes of M. tuberculosis DHFR with NADP and different inhibitors have been determined, as well as the binary complex with NADP, with resolutions ranging from 1.7 to 2.0 A. The three DHFR inhibitors are the anticancer drug methotrexate, the antimicrobial trimethoprim and Br-WR99210, an analogue of the antimalarial agent WR99210. Structural comparison of these complexes with human dihydrofolate reductase indicates that the overall protein folds are similar, despite only 26 % sequence identity, but that the environments of both NADP and of the inhibitors contain interesting differences between the enzymes from host and pathogen. Specifically, residues Ala101 and Leu102 near the N6 of NADP are distinctly more hydrophobic in the M. tuberculosis than in the human enzyme. Another striking difference occurs in a region near atoms N1 and N8 of methotrexate, which is also near atom N1 of trimethoprim, and near the N1 and two methyl groups of Br-WR99210. A glycerol molecule binds here in a pocket of the M. tuberculosis DHFR:MTX complex, while this pocket is essentially filled with hydrophobic side-chains in the human enzyme. These differences between the enzymes from pathogen and host provide opportunities for designing new selective inhibitors of M. tuberculosis DHFR.  相似文献   

7.
The structure of mouse L1210 dihydrofolate reductase (DHFR) complexed with NADPH and trimethoprim has been refined at 2.0 A resolution. The analogous complex with NADPH and methotrexate has been refined at 2.5 A resolution. These structures reveal for the first time details of drug interactions with a mammalian DHFR, which are compared with those observed from previous X-ray investigations of DHFR/inhibitor complexes. The refined L1210 structure has been used as the basis for the construction of a model of the human enzyme. There are only twenty-one sequence differences between mouse L1210 and human DHFRs, and all but two of these are located close to the molecular surface: a strong indication that the active sites are essentially identical in these two mammalian enzymes.  相似文献   

8.
Although antifolates such as trimethoprim are used in the clinical treatment of Stenotrophomonas maltophilia infection, the dihydrofolate reductase (DHFR) of this microorganism is scarcely known because it has never been isolated. Here, we describe the purification of this enzyme and kinetically characterize its inhibition by methotrexate (MTX). Upon MTX treatment, time-dependent, slow-binding inhibition was observed due to the generation of a long-lived, slowly dissociating enzyme-NADPH-inhibitor complex. Kinetic analysis revealed a one-step inhibition mechanism (K(I) = 28.9 +/- 1.9 pM) with an association rate constant (k(i)) of 3.8 x 10(7) M(-1)s(-1). Possible mechanisms for MTX binding to S. maltophilia DHFR are discussed.  相似文献   

9.
E E Howell  C Booth  M Farnum  J Kraut  M S Warren 《Biochemistry》1990,29(37):8561-8569
The adaptability of Escherichia coli dihydrofolate reductase (DHFR) is being explored by identifying second-site mutations that can partially suppress the deleterious effect associated with removal of the active-site proton donor aspartic acid-27. The Asp27----serine mutant DHFR (D27S) was previously characterized and the catalytic activity found to be greatly decreased at pH 7.0 [Howell et al. (1986) Science 231, 1123-1128]. Using resistance to trimethoprim (a DHFR inhibitor) in a genetic selection procedure, we have isolated a double-mutant DHFR gene containing Asp27----Ser and Phe137----Ser mutations (D27S+F137S). The presence of the F137S mutation increases kcat approximately 3-fold and decreases Km(DHF) approximately 2-fold over D27S DHFR values. The overall effect on kcat/Km(DHF) is a 7-fold increase. The D27S+F137S double-mutant DHFR is still 500-fold less active than wild-type DHFR at pH 7. Surprisingly, Phe137 is approximately 15 A from residue 27 in the active site and is part of a beta-bulge. We propose the F137S mutation likely causes its catalytic effect by slightly altering the conformation of D27S DHFR. This supposition is supported by the observation that the F137S mutation does not have the same kinetic effect when introduced into the wild-type and D27S DHFRs, by the altered distribution of two conformers of free enzyme [see Dunn et al. (1990)] and by a preliminary difference Fourier map comparing the D27S and D27S+F137S DHFR crystal structures.  相似文献   

10.
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   

11.
West FW  Seo HS  Bradrick TD  Howell EE 《Biochemistry》2000,39(13):3678-3689
R67 dihydrofolate reductase (DHFR) is an R-plasmid-encoded enzyme that confers clinical resistance to the antibacterial drug trimethoprim. This enzyme shows no sequence or structural homology to the chromosomal DHFRs. The active form of the protein is a homotetramer possessing D(2) symmetry and a single active-site pore. Two tryptophans occur per monomer: W38 and its symmetry-related residues (W138, W238, and W338) occur at the dimer-dimer interfaces, while W45 and its symmetry-related partners (W145, W245, and W345) occur at the monomer-monomer interfaces. Two single-tryptophan mutant genes were constructed to determine the structural and functional consequences of four mutations per tetramer. The W45F mutant retains full enzyme activity and the fluorescence environment of the unmutated W38 residues clearly monitors ligand binding and a pH dependent tetramer right harpoon over left harpoon 2 dimers equilibrium. In contrast, four simultaneous W38F mutations at the dimer-dimer interfaces result in tetramer destabilization. The ensuing dimer is relatively inactive, as is dimeric wild-type R67 DHFR. A comparison of emission spectra indicates the fluorescent signal of wild-type R67 DHFR is dominated by the contribution from W38. Equilibrium unfolding/folding curves at pH 5.0, where all protein variants are dimeric, indicate the environment monitored by the W38 residue is slightly less stable than the environment monitored by the W45 residue.  相似文献   

12.
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   

13.
Dihydrofolate reductase (DHFR) is the subject of intensive investigation since it appears to be the primary target enzyme for "antifolate" drugs, such as methotrexate and trimethoprim. Fluorescence quenching and stopped-flow fluorimetry show that the ester bond-containing tea polyphenols (-)-epigallocatechin gallate (EGCG) and (-)-epicatechin gallate (ECG) are potent and specific inhibitors of DHFR with inhibition constants (K(I)) of 120 and 82 nM, respectively. Both tea compounds showed the characteristics of slow-binding inhibitors of bovine liver DHFR. In this work, we have determined a complete kinetic scheme to explain the slow-binding inhibition and the pH effects observed during the inhibition of bovine liver DHFR by these tea polyphenols. Experimental data, based on fluorimetric titrations, and transient phase and steady-state kinetic studies confirm that EGCG and ECG are competitive inhibitors with respect to 7,8-dihydrofolate, which bind preferentially to the free form of the enzyme. The origin of their slow-binding inhibition is proposed to be the formation of a slow dissociation ternary complex by the reaction of NADPH with the enzyme-inhibitor complex. The pH controls both the ionization of critical catalytic residues of the enzyme and the protonation state of the inhibitors. At acidic pH, EGCG and ECG are mainly present as protonated species, whereas near neutrality, they evolve toward deprotonated species due to ionization of the ester-bonded gallate moiety (pK = 7.8). Although DHFR exhibits different affinities for the protonated and deprotonated forms of EGCG and ECG, it appears that the ionization state of Glu-30 in DHFR is critical for its inhibition. The physiological implications of these pH dependencies are also discussed.  相似文献   

14.
Dihydrofolate reductase (DHFR; EC 1.5.1.3) was purified to homogeneity from soybean seedlings by affinity chromatography on methotrexate-aminohexyl Sepharose, gel filtration on Ultrogel AcA-54, and Blue Sepharose chromatography. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the enzyme gave a single protein band corresponding to a molecular weight of 22,000. The enzyme is not a 140,000 Da heteropolymer as reported by others. Amino acid sequence-specific antibodies to intact human DHFR and also antibodies to CNBr-generated fragments of human DHFR bound to the plant enzyme on Western blots and cross-reacted significantly in immunoassays, indicating the presence of sequence homology between the two enzymes. The plant and human enzymes migrated similarly on nondenaturing polyacrylamide electrophoretic gels as monitored by activity staining with a tetrazolium dye. The specific activity of the plant enzyme was 15 units/mg protein, with a pH optimum of 7.4. Km values of the enzyme for dihydrofolate and NADPH were 17 and 30 microM, respectively. Unlike other eukaryotic enzymes, the plant enzyme showed no activation with organic mercurials and was inhibited by urea and KCl. The affinity of the enzyme for folate was relatively low (I50 = 130 microM) while methotrexate bound very tightly (KD less than 10(-10) M). Binding of pyrimethamine to the plant enzyme was weaker, while trimethoprim binding was stronger than to vertebrate DHFR. Trimetrexate, a very potent inhibitor of the human and bacterial enzymes showed weak binding to the plant enzyme. However, certain 2,4-diaminoquinazoline derivatives were very potent inhibitors of the plant DHFR. Thus, the plant DHFR, while showing similarity to the vertebrate and bacterial enzymes in terms of molecular weight and immunological cross-reactivity, can be distinguished from them by its kinetic properties and interaction with organic mercurials, urea, KCl and several antifolates.  相似文献   

15.
16.
We overexpressed and purified from Escherichia coli the dihydrofolate reductase (DHFR) of the gammaherpesviruses human herpesvirus 8 (HHV-8), herpesvirus saimiri (HVS), and rhesus rhadinovirus (RRV). All three enzymes proved catalytically active. The K(m) value of HHV-8 DHFR for dihydrofolate (DHF) was 2.02+/-0.44 microM, that of HVS DHFR was 4.31+/-0.56 microM, and that of RRV DHFR is 7.09+/-0.11 microM. These values are approximately 5-15-fold higher than the K(m) value reported for the human DHFR. The K(m) value of HHV-8 DHFR for NADPH was 1.31+/-0.23 microM, that of HVS DHFR was 3.78+/-0.61 microM, and that of RRV DHFR was 7.47+/-0.59 microM. These values are similar or slightly higher than the corresponding K(m) value of the human enzyme. Methotrexate, aminopterin, trimethoprim, pyrimethamine, and N(alpha)-(4-amino-4-deoxypteroyl)-N(delta)-hemiphthaloyl-L-ornithine (PT523), all well-known folate antagonists, inhibited the DHFR activity of the three gammaherpesviruses competitively with respect to DHF but proved markedly less inhibitory to the viral than towards the human enzyme.  相似文献   

17.
Coxiella burnetii is a gram-negative bacterium able to infect several eukaryotic cells, mainly monocytes and macrophages. It is found widely in nature with ticks, birds, and mammals as major hosts. C. burnetii is also the biological warfare agent that causes Q fever, a disease that has no vaccine or proven chemotherapy available. Considering the current geopolitical context, this fact reinforces the need for discovering new treatments and molecular targets for drug design against C. burnetii. Among the main molecular targets against bacterial diseases reported, the enzyme dihydrofolate reductase (DHFR) has been investigated for several infectious diseases. In the present work, we applied molecular modeling techniques to evaluate the interactions of known DHFR inhibitors in the active sites of human and C. burnetii DHFR (HssDHFR and CbDHFR) in order to investigate their potential as selective inhibitors of CbDHFR. Results showed that most of the ligands studied compete for the binding site of the substrate more effectively than the reference drug trimethoprim. Also the most promising compounds were proposed as leads for the drug design of potential CbDHFR inhibitors.  相似文献   

18.
R67 dihydrofolate reductase (DHFR) catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate using NADPH as a cofactor. This enzyme is a homotetramer possessing 222 symmetry, and a single active site pore traverses the length of the protein. A promiscuous binding surface can accommodate either DHF or NADPH, thus two nonproductive complexes can form (2NADPH or 2DHF) as well as a productive complex (NADPH.DHF). The role of water in binding was monitored using a number of different osmolytes. From isothermal titration calorimetry (ITC) studies, binding of NADPH is accompanied by the net release of 38 water molecules. In contrast, from both steady state kinetics and ITC studies, binding of DHF is accompanied by the net uptake of water. Although different osmolytes have similar effects on NADPH binding, variable results are observed when DHF binding is probed. Sensitivity to water activity can also be probed by an in vivo selection using the antibacterial drug, trimethoprim, where the water content of the media is decreased by increasing concentrations of sorbitol. The ability of wild type and mutant clones of R67 DHFR to allow host Escherichia coli to grow in the presence of trimethoprim plus added sorbitol parallels the catalytic efficiency of the DHFR clones, indicating water content strongly correlates with the in vivo function of R67 DHFR.  相似文献   

19.
Oligonucleotide-directed, site-specific mutagenesis was used to convert phenylalanine-31 of human recombinant dihydrofolate reductase (DHFR) to leucine. This substitution was of interest in view of earlier chemical modification studies (Kumar et al., 1981) and structural studies based on X-ray crystallographic data (Matthews et al., 1985a,b) which had implicated the corresponding residue in chicken liver DHFR, Tyr-31, in the binding of dihydrofolate. Furthermore, this particular substitution allowed testing of the significance of protein sequence differences between mammalian and bacterial reductases at this position with regard to the species selectivity of trimethoprim. Both wild-type (WT) and mutant (F31L) enzymes were expressed and purified by using a heterologous expression system previously described (Prendergast et al., 1988). Values of the inhibition constants (Ki values) for trimethoprim were 1.00 and 1.08 microM for WT and F31L, respectively. Thus, the presence of phenylalanine at position 31 in human dihydrofolate reductase does not contribute to the species selectivity of trimethoprim. The Km values for nicotinamide adenine dinucleotide phosphate (reduced) (NADPH) and dihydrofolate were elevated 10.8-fold and 9.4-fold, respectively, for the mutant enzyme, whereas the Vmax increased only 1.8-fold. Equilibrium dissociation constants (KD values) were obtained for the binding of NADPH and dihydrofolate in binary complexes with each enzyme. The KD for NADPH is similar in both WT and F31L, whereas the KD for dihydrofolate is 43-fold lower in F31L. Values for dihydrofolate association rate constants (kon) with enzyme and enzyme-NADPH complexes were measured by stopped-flow techniques.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
The effect of mutations on the interactions between dimers in R67 dihydrofolate reductase (R67 DHFR), a tetrameric enzyme conferring resistance to trimethoprim, was investigated by site-directed mutagenesis combined with phenotypic, enzymatic, and biochemical analysis.Some 14 mutants at two positions involved in a hydrogen bond between dimers were constructed. All were shown to be dimers. However, complementation between pairs of dimeric mutated proteins resulted in the restoration of the enzymatic activity and heterotetramer formation. A combinatorial approach was set up to create efficiently such heterotetramers and identify the complementing pairs of mutations. A dozen of such pairs were found. An accurate method was set up to measure the association of the complementing dimers in a "quasi-isologous" heterotetramer and used to study the effects of mutations and pH on the association. Thus, the pair of proteins bearing respectively the S59A and H62L mutations was shown to form heterotetramers with catalytic properties close to those of the wild-type protein. Its association was as strong as that of the wild-type protein at cytoplasmic pH (6. 5), and was more stable at lower pH values.A double-mutant protein bearing simultaneously the S59A and H62L mutations was produced and analyzed. Its association was weakened by 1.2 kcal/mol as compared to the wild-type enzyme at pH 6.5 but was insensitive to pH. Comparing the energy of association between dimers in the wild-type protein, the heterotetramer and the double mutant allowed us to dissect the effects of the pH and of the molecular context on a subset of interactions between the R67 DHFR subunits.  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