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.
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   

2.
R67 dihydrofolate reductase (R67 DHFR) catalyzes the transfer of a hydride ion from NADPH to dihydrofolate, generating tetrahydrofolate. The homotetrameric enzyme provides a unique environment for catalysis as both ligands bind within a single active site pore possessing 222 symmetry. Mutation of one active site residue results in concurrent mutation of three additional symmetry-related residues, causing large effects on binding of both ligands as well as catalysis. For example, mutation of symmetry-related tyrosine 69 residues to phenylalanine (Y69F), results in large increases in Km values for both ligands and a 2-fold rise in the kcat value for the reaction (Strader, M. B., Smiley, R. D., Stinnett, L. G., VerBerkmoes, N. C., and Howell, E. E. (2001) Biochemistry 40, 11344-11352). To understand the interactions between specific Tyr-69 residues and each ligand, asymmetric Y69F mutants were generated that contain one to four Y69F mutations. A general trend observed from isothermal titration calorimetry and steady-state kinetic studies of these asymmetric mutants is that increasing the number of Y69F mutations results in an increase in the Kd and Km values. In addition, a comparison of steady-state kinetic values suggests that two Tyr-69 residues in one half of the active site pore are necessary for NADPH to exhibit a wild-type Km value. A tyrosine 69 to leucine mutant was also generated to approach the type(s) of interaction(s) occurring between Tyr-69 residues and the ligands. These studies suggest that the hydroxyl group of Tyr-69 is important for interactions with NADPH, whereas both the hydroxyl group and hydrophobic ring atoms of the Tyr-69 residues are necessary for proper interactions with dihydrofolate.  相似文献   

3.
4.
The active sites of all bacterial and vertebrate dihydrofolate reductases that have been examined have a tryptophan residue near the binding sites for NADPH and dihydrofolate. In cases where the three-dimensional structure has been determined by X-ray crystallography, this conserved tryptophan residue makes hydrophobic and van der Waals interactions with the nicotinamide moiety of bound NADPH, and its indole nitrogen interacts with the C4 oxygen of bound folate through a bridge provided by a bound water molecule. We have addressed the question of why even the very conservative replacement of this tryptophan by phenylalanine does not seem to occur naturally. Human dihydrofolate reductase with this replacement of tryptophan by phenylalanine has increased rate constants for dissociation of substrates and products and a considerably decreased rate of hydride transfer. These cause some changes in steady-state kinetic behavior, including substantial increases in Michaelis constants for NADPH and dihydrofolate, but there is also a 3-fold increase in kcat. The branched mechanistic pathway for this enzyme has been completely defined and is sufficiently different from that of wild-type enzyme to cause changes in some transient-state kinetics. The most critical changes resulting from the amino acid substitution appear to be a 50% decrease in stability and a decrease in efficiency from 69% to 21% under intracellular conditions.  相似文献   

5.
Site-directed mutagenesis was used to generate mutants of recombinant mouse dihydrofolate reductase to test the role of some amino acids in the binding of two inhibitors, methotrexate and trimethoprim. Eleven mutations changing eight amino acids at positions all involved in hydrogen bonding or hydrophobic interactions with dihydrofolate or one of the two inhibitors were tested. Nine mutants were obtained by site-directed mutagenesis and two were spontaneous mutants previously obtained by in vivo selection (Grange, T., Kunst, F., Thillet, J., Ribadeau-Dumas, B., Mousseron, S., Hung, A., Jami, J., and Pictet, R. (1984) Nucleic Acids Res. 12, 3585-3601). The choice of the mutated positions was based on the knowledge of the active site of chicken dihydrofolate reductase established by x-ray crystallographic studies since the sequences of all known eucaryotic dihydrofolate reductases are greatly conserved. Enzymes were produced in great amounts and purified using a plasmid expressing the mouse cDNA into a dihydrofolate reductase-deficient Escherichia coli strain. The functional properties of recombinant mouse dihydrofolate reductase purified from bacterial extracts were identical to those of dihydrofolate reductase isolated from eucaryotic cells. The Km(NADPH) values for all the mutants except one (Leu-22----Arg) were only slightly modified, suggesting that the mutations had only minor effects on the ternary conformation of the enzyme. In contrast, all Km(H2folate) values were increased, since the mutations were located in the dihydrofolate binding site. The catalytic activity was also modified for five mutants with, respectively, a 6-, 10-, 36-, and 60-fold decrease of Vmax for Phe-31----Arg, Ile-7----Ser, Trp 24----Arg and Leu-22----Arg mutants and a 2-fold increase for Val-115----Pro. All the mutations affected the binding of methotrexate and six, the binding of trimethoprim: Ile-7----Ser, Leu-22----Arg, Trp-24----Arg, Phe-31----Arg, Gln-35----Pro and Phe-34----Leu. The relative variation of Ki for methotrexate and trimethoprim were not comparable from one mutant to the next, reflecting the different binding modes of the two inhibitors. The mutations which yielded the greatest increases in Ki are those which involved amino acids making hydrophobic contacts with the inhibitor.  相似文献   

6.
The binding site residue Trp-24 is conserved in all vertebrate and bacterial dihydrofolate reductases of known sequence. To determine its effects on enzyme properties, a Trp-24 to Phe-24 mutant (W-24-F) of human dihydrofolate reductase has been constructed by oligodeoxynucleotide site-directed mutagenesis. The W-24-F mutant enzyme appears to have a more open or flexible conformation as compared to the wild-type human dihydrofolate reductase on the basis of results of a number of studies. These studies include competitive ELISA using peptide-specific antibodies against human dihydrofolate reductase, thermal stability, and protease susceptibility studies of both mutant W-24-F and wild-type enzymes. It is concluded that Trp-24 is important for maintaining the structural integrity of the native enzymes. Changes in relative fluorescence quantum yield indicate that Trp-24 is buried and its fluorescence quenched relative to the other two tryptophan residues in the wild-type human reductase. Kinetic studies indicate that kcat values for W-24-F are increased in the pH range of 4.5-8.5 with a 5-fold increase at pH 7.5 as compared to the wild-type enzyme. However, the catalytic efficiency of W-24-F decreases rapidly as the pH is increased from 7.5 to 9.5. The Km values for dihydrofolate are also increased for W-24-F in the pH range of 4.5-9.5 with a 30-fold increase at pH 7.5, while the Km value for NADPH increases only ca. 1.4-fold at pH 7.5 as compared to the wild type.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
We have previously described methotrexate-resistant Chinese hamster ovary cells which appear to contain normal levls of a structurally altered dihydrofolate reductase (EC 1.5.1.3) (Flintoff, W.F., Davidson, S.V., and Siminovitch, L. (1976) Somatic Cell Genet.2,245-261). By selecting for increased resistance form these class I cells, class III resistant cells were isolated which appeared to possess an increased activity of the altered enzyme. In the report, we describe the purification and several properties of the reductase from wild-type cells, two independently selected class I cells, and class III resistant cell. The reductases from wild-type and resistant cells had similar specific activities using folate and dihydrofolate as substrates, and similar molecular weights as determined by sodium dodecyl sulfate gel electrophoresis. The mutant enzymes, however, were about six- to eight-fold more resistant to inhibition by methotrexate than the wild-type enzyme, suggesting a decreased affinity of the mutant reductases to methotrexate-binding. Small differences between various enzymes were also seen in other physicochemical properties such as pH optima and Km values for folate, and in their heat stabilities, which suggest that different structural alterations may lead to the same mutant phenotype. As expected from earlier studies with crude extracts, class III cells did produce a higher (about 10-fold) yield of the reductase than the class I or wild-type cells.  相似文献   

8.
We have applied site-directed mutagenesis methods to change the conserved tryptophan-22 in the substrate binding site of Escherichia coli dihydrofolate reductase to phenylalanine (W22F) and histidine (W22H). The crystal structure of the W22F mutant in a binary complex with the inhibitor methotrexate has been refined at 1.9-A resolution. The W22F difference Fourier map and least-squares refinement show that structural effects of the mutation are confined to the immediate vicinity of position 22 and include an unanticipated 0.4-A movement of the methionine-20 side chain. A conserved bound water-403, suspected to play a role in the protonation of substrate DHF, has not been displaced by the mutation despite the loss of a hydrogen bond with tryptophan-22. Steady-state kinetics, stopped-flow kinetics, and primary isotope effects indicate that both mutations increase the rate of product tetrahydrofolate release, the rate-limiting step in the case of the wild-type enzyme, while slowing the rate of hydride transfer to the point where it now becomes at least partially rate determining. Steady-state kinetics show that below pH 6.8, kcat is elevated by up to 5-fold in the W22F mutant as compared with the wild-type enzyme, although kcat/Km(dihydrofolate) is lower throughout the observed pH range. For the W22H mutant, both kcat and kcat/Km(dihydrofolate) are substantially lower than the corresponding wild-type values. While both mutations weaken dihydrofolate binding, cofactor NADPH binding is not significantly altered. Fitting of the kinetic pH profiles to a general protonation scheme suggests that the proton affinity of dihydrofolate may be enhanced upon binding to the enzyme. We suggest that the function of tryptophan-22 may be to properly position the side chain of methionine-20 with respect to N5 of the substrate dihydrofolate.  相似文献   

9.
Lysine-54 of human dihydrofolate reductase (hDHFR) appears to be involved in the interaction with the 2'-phosphate of NADPH and is conserved as a basic residue in other species. Studies have suggested that in Lactobacillus casei dihydrofolate reductase Arg-43, the homologous residue at this position, plays an important role in the binding of NADPH and in the differentiation of Km values for NADPH and NADH. A Lys-54 to Gln-54 mutant (K54Q) of hDHFR has been constructed by oligodeoxynucleotide-directed mutagenesis in order to study the role of Lys-54 in differentiating Km and Kcat values for NADPH and NADH as well as in other functions of hDHFR. The purpose of this paper is to delineate in quantitative terms the magnitude of the effect of the Lys-54 to Gln-54 replacement on the various kinetic parameters of hDHFR. Such quantitative effects cannot be predicted solely on the basis of X-ray structures. The Km for NADPH for the K54Q mutant enzyme is 58-fold higher, while the Km for NADH for K54Q is only 3.9-fold higher than that of the wild type, indicating that the substitution of Lys-54 with Gln-54 decreases the apparent affinity of the enzyme for NADPH dramatically, but has a lesser effect on the apparent affinity for NADH.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The role of the active site residue phenylalanine-31 (Phe31) for recombinant human dihydrofolate reductase (rHDHFR) has been probed by comparing the kinetic behavior of wild-type enzyme (wt) with mutant in which Phe31 is replaced by leucine (F31L rHDHFR). At pH 7.65 the steady-state kcat is almost doubled, but the rate constant for hydride transfer is decreased to less than half that for wt enzyme, as is the rate of the obligatory isomerization of the substrate complex that precedes hydride transfer. Although steady-state measurements indicated that the mutation causes large increases in Km for both substrates, dissociation constants for many complexes are decreased. These apparent paradoxes are due to major mutation-induced decreases in rate constants (koff) for dissociation of folate, dihydrofolate, and tetrahydrofolate from all of their complexes. This results in a mechanism proceeding almost entirely by only one of the two pathways used by wt enzyme. Other consequences of these changes are a much altered dependence of steady-state kcat on pH, inhibition rather than activation by tetrahydrofolate, absence of hysteresis in transient-state kinetics, and a decrease in enzyme efficiency under physiological conditions. The results indicate that there is no quantitative correlation between dihydrofolate binding and the rate of hydride transfer for this enzyme.  相似文献   

11.
D J Murphy  S J Benkovic 《Biochemistry》1989,28(7):3025-3031
The strictly conserved residue leucine-54 of Escherichia coli dihydrofolate reductase forms part of the hydrophobic wall which binds the p-aminobenzoyl side chain of dihydrofolate. In addition to the previously reported glycine-54 mutant, isoleucine-54 and asparagine-54 substitutions have been constructed and characterized with regard to their effects on binding and catalysis. NADP+ and NADPH binding is virtually unaffected with the exception of a 15-fold decrease in NADPH dissociation from the Gly-54 mutant. The synergistic effect of NADPH on tetrahydrofolate dissociation seen in the wild-type enzyme is lost in the isoleucine-54 mutant: little acceleration is seen in tetrahydrofolate dissociation when cofactor is bound, and there is no discrimination between reduced and oxidized cofactor. The dissociation constants for dihydrofolate and methotrexate increase in the order Leu less than Ile less than Asn less than Gly, varying by a maximum factor of 1700 for dihydrofolate and 6300 for methotrexate. Despite these large changes in binding affinity, the hydride transfer rate of 950 s-1 in the wild-type enzyme is decreased by a constant factor of ca. 30 (2 kcal/mol) regardless of the mutant. Thus, the contributions of residue 54 to binding and catalysis appear to have been separated.  相似文献   

12.
Methotrexate-resistant forms of human dihydrofolate reductase have the potential to protect healthy cells from the toxicity of methotrexate (MTX), to improve prognosis during cancer therapy. It has been shown that synergistic MTX-resistance can be obtained by combining two active-site mutations that independently confer weak MTX-resistance. In order to obtain more highly MTX-resistant human dihydrofolate reductase (hDHFR) variants for this application, we used a semi-rational approach to obtain combinatorial active-site mutants of hDHFR that are highly resistant towards MTX. We created a combinatorial mutant library encoding various amino acids at residues Phe31, Phe34 and Gln35. In vivo library selection was achieved in a bacterial system on medium containing high concentrations of MTX. We characterized ten novel MTX-resistant mutants with different amino acid combinations at residues 31, 34 and 35. Kinetic and inhibition parameters of the purified mutants revealed that higher MTX-resistance roughly correlated with a greater number of mutations, the most highly-resistant mutants containing three active site mutations (Ki(MTX)=59-180 nM; wild-type Ki(MTX)<0.03 nM). An inverse correlation was observed between resistance and catalytic efficiency, which decreased mostly as a result of increased KM toward the substrate dihydrofolate. We verified that the MTX-resistant hDHFRs can protect eukaryotic cells from MTX toxicity by transfecting the most resistant mutants into DHFR-knock-out CHO cells. The transfected variants conferred survival at concentrations of MTX between 100-fold and >4000-fold higher than the wild-type enzyme, the most resistant triple mutant offering protection beyond the maximal concentration of MTX that could be included in the medium. These highly resistant variants of hDHFR offer potential for myeloprotection during administration of MTX in cancer treatment.  相似文献   

13.
Resistance to the antifolates methasquin and amethopterin has been studied in new strains of Streptococcus faecium var. durans. Two methasquin-resistant strains (SF/MQ, SF/MQ(T)) and an amethopterin-resistant strain (SF/AM) were selected independently from the wild-type S. faecium var. durans (SF/O). SF/MQ(T) is a thymine auxotroph. Total dihydrofolate reductase activity was elevated in each of the resistant strains. The greatest increase (36-fold) was observed in extracts of SF/AM. The methasquin-resistant strains, SF/MQ and SF/MQ(T), had 29-fold and 8-fold, respectively, more dihydrofolate reductase activity than the parental strain. Total dihydrofolate reductase activity of SF/O was separable by gel filtration into two components: a folate reductase (11%) and a specific dihydrofolate reductase (89%). Folate reductase activity was associated with 88% of the total dihydrofolate reductase activity of SF/MQ(T), with specific dihydrofolate reductase activity accounting for the remaining 12%. In SF/MQ and SF/AM, folate reductase activity was associated with 97% of the total dihydrofolate reductase activity. Studies of the inhibition by methasquin and amethopterin of partially purified folate reductase and specific dihydrofolate reductase of the mutant strains suggested that resistance was not accompanied by changes in the affinities of these enzymes for either antifolate.  相似文献   

14.
We describe the development of resistance to trimetrexate and piritrexim (BW 301U) by a stepwise selection protocol in Chinese hamster ovary cells. Selection in trimetrexate resulted in initial resistance as a result of dihydrofolate reductase gene amplification. Several trimetrexate-resistant variants that display 250-340-fold and 25-50-fold resistance to lipophilic and hydrophilic antifolates, respectively, were established. Increased antifolate resistance was associated with a prominent overexpression of dihydrofolate reductase as determined from the elevated folate reductase activity, cellular labeling with fluorescein-methotrexate, and steady-state mRNA levels as a result of a consistent dihydrofolate reductase gene amplification. However, upon subsequent incremental increases in trimetrexate, further resistance was also associated with amplification of the multidrug resistance gene. This resulted in overexpression of P-glycoprotein and a subsequent 20-50-fold collateral resistance to pleiotropic drugs such as adriamycin, actinomycin D, vinca alkaloids, etoposide, and colchicine. In contrast, initial resistance following selection with low piritrexim concentrations resulted from an unknown mechanism(s) not involving overproduction of either dihydrofolate reductase or P-glycoprotein. This piritrexim resistance was shared with trimetrexate but not with methotrexate. Upon further selection with piritrexim, resistant variants emerge with amplified dihydrofolate reductase but not with multidrug resistance genes. These variants were subsequently resistant to both hydrophilic and lipophilic folate antagonists but retained sensitivity to pleiotropic drugs. The pattern of resistance with methotrexate, trimetrexate, and piritrexim shared a common mechanism, dihydrofolate reductase gene amplification, but differed regarding the additional amplification of the multidrug resistance gene in trimetrexate-resistant cells as well as the emergence of an additional unknown mechanism(s) of resistance to lipid-soluble antifolates upon initial selection in piritrexim.  相似文献   

15.
Pediococcus cerevisiae/AMr, resistant to amethopterin, possesses a higher dihydrofolate reductase (5, 6, 7, 8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) activity than the parent, a folate-permeable and thus amethopterin-susceptible strain and than the wild-type. The properties of dihydrofolate reductase from the three strains have been compared. Temperature, pH optima, heat stability, as well amethopterin binding did not reveal significant differences between the enzymes from the susceptible and resistant strains. The enzyme from the wild-type was 10 times more sensitive to inhibition by amethopterin and more susceptible to heat denaturation. The apparent Km values for dihydrofolate in enzymes from the three strains were in the range of 4.8--7.2 muM and for NADPH 6.5--8.0 muM. The amethopterin-resistant strain exhibited cross-resistance to trimethoprim and was about 40-fold more resistant to the latter than the sensitive parent and the wild-type. The resistance to trimethoprim appears to be a direct result of the increased dihydrofolate reductase activity. Inhibition of dihydrofolate reductase activity by this drug was similar in the three strains. 10--20 nmol caused 50% inhibition of 0.02 enzyme unit. Trimethoprim was about 10 000 times less effective inhibitor of dihydrofolate reductase than amethopterin. The cell extract of the AMr strain possessed a folate reductase activity three times higher than that of the sensitive strain. The activities of other folate-related enzymes like thymidylate synthetase and 10-formyltetrahydrofolate synthetase (formate: tetrahydrofolate ligase (ADP-forming), EC 6.3.4.3) were similar in the three strains studied.  相似文献   

16.
Initial and steady-state uptakes of serine and phenylalanine by human fibroblasts and human colon tumour cells were studied applying a double isotope dilution technique to perfused populations of cultivated cells retained on microcarrier beads. This new method permits the differentiation of the unidirectional transport parameters and can also distinguish between membrane-associated processes and independently intracellular events in isolated cells. High initial L-serine uptake values in colon adenocarcinoma cells became negative under steady-state conditions. To determine if the observed negative L-serine uptake was produced by the rapid efflux of intracellular L-[3H]serine, the cells were treated with methotrexate (MTX) (an inhibitor of cytosolic dihydrofolate reductase). The modified curve of L-[3H]serine uptake after MTX treatment suggests that, under these experimental conditions, net serine transport is non concentrative in colon tumour cells and could be modulated by the rate of intracellular serine metabolism; it also suggests that MTX does not directly affect serine transport in perfused human colon adenocarcinoma cells. Initial and steady-state uptakes of phenylalanine were high in both fibroblasts and tumour cells and were unaffected by MTX treatment.  相似文献   

17.
Adapting metabolic enzymes of microorganisms to low temperature environments may require a difficult compromise between velocity and affinity. We have investigated catalytic efficiency in a key metabolic enzyme (dihydrofolate reductase) of Moritella profunda sp. nov., a strictly psychrophilic bacterium with a maximal growth rate at 2 degrees C or less. The enzyme is monomeric (Mr=18,291), 55% identical to its Escherichia coli counterpart, and displays Tm and denaturation enthalpy changes much lower than E. coli and Thermotoga maritima homologues. Its stability curve indicates a maximum stability above the temperature range of the organism, and predicts cold denaturation below 0 degrees C. At mesophilic temperatures the apparent Km value for dihydrofolate is 50- to 80-fold higher than for E. coli, Lactobacillus casei, and T. maritima dihydrofolate reductases, whereas the apparent Km value for NADPH, though higher, remains in the same order of magnitude. At 5 degrees C these values are not significantly modified. The enzyme is also much less sensitive than its E. coli counterpart to the inhibitors methotrexate and trimethoprim. The catalytic efficiency (kcat/Km) with respect to dihydrofolate is thus much lower than in the other three bacteria. The higher affinity for NADPH could have been maintained by selection since NADPH assists the release of the product tetrahydrofolate. Dihydrofolate reductase adaptation to low temperature thus appears to have entailed a pronounced trade-off between affinity and catalytic velocity. The kinetic features of this psychrophilic protein suggest that enzyme adaptation to low temperature may be constrained by natural limits to optimization of catalytic efficiency.  相似文献   

18.
L Y Li  S J Benkovic 《Biochemistry》1991,30(6):1470-1478
The alpha C-helix of Escherichia coli dihydrofolate reductase has been converted to its counterpart in Lactobacillus casei by a triple mutation in the helix (H45R, W47Y, and I50F). These changes result in a 2-fold increase in the steady-state reaction rate (kcat = 26 s-1) that is limited by an increased off rate for the release of tetrahydrofolate (koff = 40 s-1 versus 12 s-1). On the other hand the mutant protein exhibits a 10-fold increase in the KM value (6.8 microM) for dihydrofolate and a 10-fold decrease in the rate of hydride transfer (85 s-1) from NADPH to dihydrofolate. The elevated rate of tetrahydrofolate release upon the rebinding of NADPH, a characteristic of the wild-type enzyme-catalyzed reaction, is diminished. The intrinsic pKa (6.4) of the mutant enzyme binary complex with NADPH is similar to that of the wild type, but the pKa of the ternary complex is increased to 7.3, about on pH unit higher than the wild-type value. Further mutagenesis (G51P and an insertion of K52) was conducted to incorporate a hairpin turn unique to the C-terminus of the alpha C-helix of the L. casei enzyme in order to adjust a possible dislocation of the new helix. The resultant pentamutant enzyme shows restoration of many of the kinetic parameters, such as kcat (12 s-1), KM (1.1 microM for dihydrofolate), and khyd (526 s-1), to the wild-type values. The synergism in the product release is also largely restored. A substrate-induced conformational change responsible for the fine tuning of the catalytic process was found to be associated with the newly installed hairpin structure. The Asp27 residue of the mutant enzyme was found to be reprotonated before tetrahydrofolate release.  相似文献   

19.
Arginine-70 of human dihydrofolate reductase (hDHFR) is a highly conserved residue which X-ray crystallographic data have shown to interact with the alpha-carboxylate of the terminal L-glutamate moiety of either folic acid or methotrexate (MTX). The rationale for this study was to introduce a conservative amino acid residue change at position 70 (Arg----Lys) which might function as a titratable group and, thus, reveal possible quantitative changes in ligand binding and kinetic parameters as a function of pH. Such a mutant enzyme (R70K) has been constructed and expressed by using site-directed mutagenesis techniques. This substitution has a dramatic effect on the binding of MTX, which displays a 22,600-fold increase in the dissociation constant (KD) at pH 7.5 compared to that of the reported wild-type enzyme value. At this pH, the KD value for dihydrofolate (FAH2) for the R70K enzyme shows only a 7-fold increase over that for the wild-type hDHFR. The pH profiles of the Michaelis and dissociation constants for FAH2 and KD values for MTX for the mutant enzyme all show a 7-8-fold increase from pH 7.5 to 8.5 as compared to its wild-type counterpart. The binding of NADPH or the nonclassical inhibitor trimetrexate (TMQ) to either the wild-type or the mutant enzyme does not show such pH-dependent characteristics. Thus, since FAH2 and MTX interact with the guanidinium side chain of arginine-70 in the wild-type hDHFR, the replacement of this residue with a lysine in the R70K mutant appears to have resulted in the introduction of a titratable group with a perturbed pKa value of ca. 8.3.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
We have explored the substrate protonation mechanism of Escherichia coli dihydrofolate reductase by changing the location of the proton donor. A double mutant was constructed in which the proton donor of the wild-type enzyme, aspartic acid-27, has been changed to serine and simultaneously an alternative proton donor, glutamic acid, has replaced threonine at position 113. The active site of the resulting variant enzyme molecule should therefore somewhat resemble that proposed for the R67 plasmid-encoded dihydrofolate reductase [Matthews, D. A., Smith, S. L., Baccanari, D. P., Burchall, J. J., Oatley, S. J., & Kraut, J. (1986) Biochemistry 25, 4194]. At pH 7, the double-mutant enzyme has a 3-fold greater kcat and an unchanged Km(dihydrofolate) as compared with the single-mutant Asp-27----Ser enzyme described previously [Howell, E. E., Villafranca, J. E., Warren, M. S., Oatley, S. J., & Kraut, J. (1986) Science (Washington, D.C.) 231, 1123]. Additionally, its activity vs pH profiles together with observed deuterium isotope effects, suggest that catalysis depends on an acidic group with a pKa of 8. It is concluded that the dihydropteridine ring of a bound substrate molecule can indeed be protonated by a glutamic acid side chain at position 113 (instead of an aspartic acid side chain at position 27), but with greatly decreased efficiency: at pH 7, the double mutant still has a 25-fold lower kcat (1.2 s-1) and a 2900-fold lower kcat/km(dihydrofolate) (8.6 X 10(3) s-1 M-1) than the wild-type enzyme.  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