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.
The binding of trimethoprim to dihydrofolate reductase from L1210 mouse lymphoma cells has been studied by measuring the changes in chemical shift of nuclei of the ligand that accompanying binding. The 6- and 2',6'-proton chemical shifts of bound trimethoprim have been determined by transfer of saturation experiments, and the 2-carbon chemical shift has been determined by using [2-13C]trimethoprim. The changes in proton chemical shift are substantially smaller than those accompanying binding to bacterial dihydrofolate reductase [Cayley, P. J., Albrand, J. P., Feeney, J., Robert, G. C. K., Piper, E. A., & Burgen, A. S. V. (1979) Biochemistry 18, 3886]. It is shown that this difference arises largely from the fact that trimethoprim adopts different conformations when bound to mammalian and to bacterial dihydrofolate reductase. The proton chemical shifts are interpreted in terms of ring-current contributions from the two aromatic rings of trimethoprim itself and the nearby aromatic amino acid residues of the enzyme. The latter have been located by using the refined crystallographic coordinates of the Lactobacillus casei and Escherichia coli reductases in their complexes with methotrexate [Bolin, J. T., Filman, D. J., Matthews, D. A. & Kraut, J. (1982) J. Biol. Chem. 257, 13650], under the assumption that, as indicated by the 13C chemical shifts, the diaminopyrimidine ring of trimethoprim binds in the same way as does the corresponding part of methotrexate. With use of these assumptions, the conformation of trimethoprim bound to the dihydrofolate reductases from L. casei, E. coli, and L1210 cells has been calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Circular-dichroism spectra (200--450 nm) were recorded for Lactobacillus casei MTX/R dihydrofolate reductase and its complexes with substrates, inhibitors and coenzymes. These spectra are compared with those reported by others for dihydrofolate reductase from other sources. The binding of NADP+ or NADPH is associated with the perturbation of one or more aromatic amino acid residues, and there is marked enhancement of the negative c.d. band at 340 nm arising from the dihydronicotinamide chromophore of NADPH. The substrates folate and dihydrofolate give rise to substantial extrinsic c.d. bands on binding, which show a number of specific differences between enzymes from different sources. The binary complexes between the enzyme and the inhibitors methotrexate or trimethoprim also show strong c.d. bands, and these are qualitatively very similar for all dihydrofolate reductases studied so far. The ternary complexes between enzyme, NADPH and trimethoprim or methotrexate are very different from the sum of the spectra of the binary complexes. Trimethoprim leads to the disappearance of the 340 nm c.d. band of bound NADPH, whereas in the methotrexate--NADPH--enzyme ternary complex a "couplet" c.d. spectrum is observed at long wavelengths. Analysis of this latter feature suggests that it arises from a direct interaction between the dihydronicotinamide and pteridine rings in the ternary complex.  相似文献   

3.
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   

4.
Dihydrofolate reductase. The stereochemistry of inhibitor selectivity   总被引:7,自引:0,他引:7  
X-ray structural results are reported for 10 triazine and pyrimidine inhibitors of dihydrofolate reductase, each one studied as a ternary complex with NADPH and chicken dihydrofolate reductase. Analysis of these data and comparison with structural results from the preceding paper (Matthews, D.A., Bolin, J.T., Burridge, J.M., Filman, D.J., Volz, K.W., Kaufman, B. T., Beddell, C.R., Champness, J.N., Stammers, D.K., and Kraut, J. (1985) J. Biol. Chem. 260, 381-391) in which we contrasted binding of the antibiotic trimethoprim (TMP) to chicken dihydrofolate reductase on the one hand with its binding to Escherichia coli dihydrofolate reductase on the other, permit identification of differences that are important in accounting for TMP's selectivity. The crystallographic evidence strongly suggests that loss of a potential hydrogen bond between the 4-amino group of TMP and the backbone carbonyl of Val-115 when TMP binds to chicken dihydrofolate reductase but not when it binds to the E. coli reductase is the major factor responsible for this drug's more potent inhibition of bacterial dihydrofolate reductase. A key finding of the current study which is important in understanding why TMP binds differently to chicken and E. coli dihydrofolate reductases is that residues on opposite sides of the active-site cleft in chicken dihydrofolate reductase are about 1.5-2.0 A further apart than are structurally equivalent residues in the E. coli enzyme.  相似文献   

5.
Dihydrofolate reductase has been purified from a trimethoprim-resistant strain of Neisseria gonorrhoeae. The enzyme showed a single component on sodium dodecyl sulfate-polyacrylamide gel electrophoresis (Mr = 18,000) and on isoelectric focusing in 5 M urea (pI = 6.8). Although gel electrophoresis under nondenaturing conditions resolved the preparation into two enzymatically active proteins (called form 1 and form 2), they were not genetically determined isozymes. Both had a similar dihydrofolate Km (2 microM), NADPH Km (10 microM), and trimethoprim Ki (20 nM), and form 2 (the slower migrating species) was shown to be generated from form 1 by the electrophoresis conditions. The complete covalent structure of the enzyme has also been determined. It is a single polypeptide composed of 162 residues and containing 4 cysteines. The gonococcal dihydrofolate reductase shares a 35% homology with the chicken liver enzyme and a 40% homology with the Escherichia coli enzyme. Most of these identities are residues that have been implicated in the binding of NADPH and methotrexate to the E. coli and Lactobacillus casei reductases.  相似文献   

6.
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   

7.
Dihydrofolate reductase (5,6,7,8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) from an amethopterin-resistant strain of Lactobacillus casei was inactivated by 2,4-pentanedione. The inactivation appears to be due to the specific interaction of 2,4-pentanedione with lysyl residues. Inactivation is concomitant with with the modification of three lysyl residues. Both NADPH and dihydrofolate protect the enzyme against inactivation, suggesting that the critical residue(s) lies at or near their binding sites. Unlike native dihydrofolate reductase, 2,4-pentanedione-modified enzyme does not form binary complexes with either NADPH, dihydrofolate or amethopterin which are stable to gel filtration. Treatment of the modified enzyme with nucleophilic reagents such as hydroxylamine, failed to promote reactivation of the enzyme. Reactivation was achieved following gel filtration at pH 6.0 and was found to be dependent on the degree to which the enzyme was inactivated.  相似文献   

8.
To test the hypothesis that protein folding mechanisms are better conserved than amino acid sequences, the mechanisms for dihydrofolate reductases (DHFR) from human (hs), Escherichia coli (ec) and Lactobacillus casei (lc) were elucidated and compared using intrinsic Trp fluorescence and fluorescence-detected 8-anilino-1-naphthalenesulfonate (ANS) binding. The development of the native state was monitored using either methotrexate (absorbance at 380 nm) or NADPH (extrinsic fluorescence) binding. All three homologs displayed complex unfolding and refolding kinetic mechanisms that involved partially folded states and multiple energy barriers. Although the pairwise sequence identities are less than 30 %, folding to the native state occurs via parallel folding channels and involves two types of on-pathway kinetic intermediates for all three homologs. The first ensemble of kinetic intermediates, detected within a few milliseconds, has significant secondary structure and exposed hydrophobic cores. The second ensemble is obligatory and has native-like side-chain packing in a hydrophobic core; however, these intermediates are unable to bind active-site ligands. The formation of the ensemble of native states occurs via three channels for hsDHFR, and four channels for lcDHFR and ecDHFR. The binding of active-site ligands (methotrexate and NADPH) accompanies the rate-limiting formation of the native ensemble. The conservation of the fast, intermediate and slow-folding events for this complex alpha/beta motif provides convincing evidence for the hypothesis that evolutionarily related proteins achieve the same fold via similar pathways.  相似文献   

9.
R-Plasmids from a number of trimethoprim-resistant Escherichia coli and Citrobacter sp. were studied after transfer to E. coli K12 hosts. Each was found to specify a dihydrofolate reductase which was resistant to trimethoprim and Methotrexate, and which could be completely separated from the host chromosomal enzyme by gel filtration. Two distinct types of R-plasmid dihydrofolate reductases were identified. Type I enzymes, typified by the R483 enzyme previously described (Sk?ld, O., and Widh, A. (1974) J. Biol. Chem. 249, 4324-4325), are synthesized in amounts severalfold higher than the chromosomal enzyme. The 50% inhibitory concentrations (I50) of trimethoprim, Methotrexate, and aminopterin are increased several thousandfold over the corresponding values for the chromosomal enzyme. Type II R-plasmid dihydrofolate reductases are synthesized in about the same amount, or less, as the chromosomal enzyme, but are practically several hundredfold higher than those for the type I enzymes. Both types of R-plasmid dihydrofolate reductase showed little difference from the chromosomal enzyme in the binding of dihydrofolate, NADPH, folic acid, and 2,4-diaminopyrimidine.  相似文献   

10.
Characterization of Candida albicans dihydrofolate reductase   总被引:3,自引:0,他引:3  
Dihydrofolate reductase from Candida albicans was purified 31,000-fold and characterized. In addition, the C. albicans dihydrofolate reductase gene was cloned into a plasmid vector and expressed in Escherichia coli, and the enzyme was purified from this source. Both preparations showed a single protein-staining band with a molecular weight of about 25,000 on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzymes were stable and had an isoelectric point of pH 7.1 on gel isoelectric focusing. Kinetic characterization showed that the enzymes from each source had similar turnover numbers (about 11,000 min-1) and Km values for NADPH and dihydrofolate of 3-4 microM. Like other eukaryotic dihydrofolate reductases, the C. albicans enzyme exhibited weak binding affinity for the antibacterial agent trimethoprim (Ki = 4 microM), but further characterization showed that the inhibitor binding profile of the yeast and mammalian enzymes differed. Methotrexate was a tight binding inhibitor of human but not C. albicans dihydrofolate reductase; the latter had a relatively high methotrexate Ki of 150 pM. The yeast and vertebrate enzymes also differed in their interactions with KCl and urea. These two agents activate vertebrate dihydrofolate reductases but inhibited the C. albicans enzyme. The sequence of the first 36 amino-terminal amino acids of the yeast enzyme was also determined. This portion of the C. albicans enzyme was more similar to human than to E. coli dihydrofolate reductases (50% and 30% identity, respectively). Some key amino acid residues in the C. albicans sequence, such as E-30 (human enzyme numbering), were "vertebrate-like" whereas others, such as I-31, were not. These results indicate that there are physical and kinetic differences between the eukaryotic mammalian and yeast enzymes.  相似文献   

11.
Neisseria gonorrhoeae dihydrofolate reductase undergoes a time-dependent, irreversible inactivation by 2,4-diamino-5-[3,5-dimethoxy-4-(p-bromoacetamidophenoxy)benzyl] pyrimidine. The kinetics of inactivation are consistent with the reversible formation of an enzyme-inhibitor complex followed by covalent binding to the enzyme. The reversible component is competitive with dihydrofolate and has an inhibitor binding constant of 10 nM. Irreversible inactivation proceeds as a pseudo first-order process with a minimum inactivation half-time of 20 min and a Ki of 28 nM. Using radiolabeled inhibitor, it was shown that approximately 1 mol of ligand was covalently bound to the enzyme/mol of methotrexate binding site when the enzyme was completely inhibited. Radiolabeled inhibitor remained associated with the enzyme following denaturation and sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Cyanogen bromide cleavage of the 14C-labeled enzyme-inhibitor complex yielded only one radioactive polypeptide, and sequence determinations showed that His-25 was modified by covalent attachment of the inhibitor. When dihydrofolate reductases from Lactobacillus casei, Streptococcus faecium, Escherichia coli, SR-1 rodent lymphoma, and chicken liver were tested with the affinity label, only the L. casei enzyme showed a time-dependent increase in inhibition. These data, along with comparisons of known amino acid sequences and x-ray crystal structures, were used to make predictions concerning the three-dimensional conformation of the gonococcal enzyme.  相似文献   

12.
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   

13.
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   

14.
Chromosomal dihydrofolate reductase from Escherichia coli catalyzes the reduction of dihydrofolate to tetrahydrofolate using NADPH as a cofactor. The thermodynamics of ligand binding were examined using an isothermal titration calorimetry approach. Using buffers with different heats of ionization, zero to a small, fractional proton release was observed for dihydrofolate binding, while a proton was released upon NADP(+) binding. The role of water in binding was additionally monitored using a number of different osmolytes. Binding of NADP(+) is accompanied by the net release of ~5-24 water molecules, with a dependence on the identity of the osmolyte. In contrast, binding of dihydrofolate is weakened in the presence of osmolytes, consistent with "water uptake". Different effects are observed depending on the identity of the osmolyte. The net uptake of water upon dihydrofolate binding was previously observed in the nonhomologous R67-encoded dihydrofolate reductase (dfrB or type II enzyme) [Chopra, S., et al. (2008) J. Biol. Chem. 283, 4690-4698]. As R67 dihydrofolate reductase possesses a nonhomologous sequence and forms a tetrameric structure with a single active site pore, the observation of weaker DHF binding in the presence of osmolytes in both enzymes implicates cosolvent effects on free dihydrofolate. Consistent with this analysis, stopped flow experiments find betaine mostly affects DHF binding via changes in k(on), while betaine mostly affects NADPH binding via changes in k(off). Finally, nonadditive enthalpy terms when binary and ternary cofactor binding events are compared suggest the presence of long-lived conformational transitions that are not included in a simple thermodynamic cycle.  相似文献   

15.
Affinity chromatography of dihydrofolate reductase   总被引:10,自引:5,他引:5       下载免费PDF全文
1. Dihydrofolate reductase was purified from Lactobacillus casei MTX/R, and studied on affinity columns containing folic acid and methotrexate. Two forms of the enzyme were interconverted by incubation with substrates. 2. Affinity columns were prepared from agarose activated with cyanogen bromide and coupled with 1,6-diaminohexane. Stable folate derivatives were covalently attached by using a carbodi-imide condensation. 3. Columns containing folic acid retarded but did not retain the enzyme. 4. Methotrexate at pH 6.0 was particularly effective for retention of the enzyme. 5. There is selective loss of one form of the enzyme during affinity chromatography in the absence of added NADPH. This loss is due to conversion into a single enzyme form on the column. 6. NADPH has a dual effect in stabilizing the enzyme and in sensitizing it to inactivation by methotrexate, particularly in the presence of glycine. 7. Protein with affinity for methotrexate, but without dihydrofolate reductase activity, may also be eluted from the columns. 8. In a single-step procedure the enzyme was purified nearly 4000-fold from mammalian skin.  相似文献   

16.
Glutathione synthetase from Escherichia coli B showed amino acid sequence homology with mammalian and bacterial dihydrofolate reductases over 40 residues, although these two enzymes are different in their reaction mechanisms and ligand requirements. The effects of ligands of dihydrofolate reductase on the reaction of E. coli B glutathione synthetase were examined to find resemblances in catalytic function to dihydrofolate reductase. The E. coli B enzyme was potently inhibited by 7,8-dihydrofolate, methotrexate, and trimethoprim. Methotrexate was studied in detail and proved to bind to an ATP binding site of the E. coli B enzyme with K1 value of 0.1 mM. The homologous portion of the amino acid sequence in dihydrofolate reductases, which corresponds to the portion coded by exon 3 of mammalian dihydrofolate reductase genes, provided a binding site of the adenosine diphosphate moiety of NADPH in the crystal structure of dihydrofolate reductase. These analyses would indicate that the homologous portion of the amino acid sequence of the E. coli B enzyme provides the ATP binding site. This report gives experimental evidence that amino acid sequences related by sequence homology conserve functional similarity even in enzymes which differ in their catalytic mechanisms.  相似文献   

17.
13C nuclear magnetic resonance (NMR) of methotrexate, trimethoprim, and pyrimethamine enriched 90% with 13C at C2 has provided a sensitive means of detecting the state of protonation of the heterocyclic rings of these inhibitors. In each case, protonation of N1 causes an upfield movement of the chemical shift of C2 by more than 6 ppm. By this method it has been shown that, at pH values up to 9.2, methotrexate is bound to bovine liver dihydrofolate reductase with N1 of the inhibitor protonated, just as in the case of the complex with reductase from Streptococcus faecium and Lactobacillus casei. Furthermore, trimethoprim bound to reductase from any of the three sources, and pyrimethamine bound to either of the bacterial reductases also have N1 protonated even at pH values up to 10. This implies that in all cases there is a strong interaction between protonated N1 of the inhibitor and the carboxylate group of the active site aspartate or glutamate. In every case pKa of the bound inhibitor is increased by several units, a finding in accord with crystallographic evidence that inhibitor bound to L. casei reductase is in a hydrophobic environment and that N1 is not hydrogen-bonded to water. It was confirmed by titration of protein fluorescence that trimethoprim has greater affinity for bacterial reductase than for vertebrate (bovine) reductase, and that this selectivity is more marked in ternary complexes in which NADPH is also bound to the active site. However, the data cited above indicate that this difference in affinities is not due to a weaker ionic interaction between protonated N1 of trimethoprim and the bovine enzyme. Instead, binding of the trimethoprim side chain to hydrophobic sites on the enzyme must provide less binding energy in the case of the mammalian enzyme.  相似文献   

18.
R67 is a Type II dihydrofolate reductase (DHFR) that catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate by facilitating the addition of a proton to N5 of DHF and the transfer of a hydride ion from NADPH to C6. Because this enzyme is a plasmid-encoded DHFR from trimethoprim-resistant bacteria, extensive studies on R67 with various methods have been performed to elucidate its reaction mechanism. Here, Raman difference measurements, conducted on the ternary complex of R67.NADP(+).DHF believed to be an accurate mimic of the productive DHFR.NADPH.DHF complex, show that the pK(a) of N5 in the complex is less than 4. This is in clear contrast to the behavior observed in Escherichia coli DHFR, a substantially more efficient enzyme, where the pK(a) of bound DHF at N5 is increased to 6.5 compared with its solution value of 2.6. A comparison of the ternary complexes in R67 and E. coli DHFRs suggests that enzymic raising of the pK(a) at N5 can significantly increase the catalytic efficiency of the hydride transfer step. However, R67 shows that even without such a strategy an effective DHFR can still be designed.  相似文献   

19.
The ionization state of aspartate 26 in Lactobacillus casei dihydrofolate reductase has been investigated by selectively labeling the enzyme with [13Cgamma] aspartic acid and measuring the 13C chemical shifts in the apo, folate-enzyme, and dihydrofolate-enzyme complexes. Our results indicate that no aspartate residue has a pKa greater than approximately 4.8 in any of the three complexes studied. The resonance of aspartate 26 in the dihydrofolate-enzyme complex has been assigned by site-directed mutagenesis; aspartate 26 is found to have a pKa value of less than 4 in this complex. Such a low pKa value makes it most unlikely that the ionization of this residue is responsible for the observed pH profile of hydride ion transfer [apparent pKa = 6.0; Andrews, J., Fierke, C. A., Birdsall, B., Ostler, G., Feeney, J., Roberts, G. C. K., and Benkovic, S. J. (1989) Biochemistry 28, 5743-5750]. Furthermore, the downfield chemical shift of the Asp 26 (13)Cgamma resonance in the dihydrofolate-enzyme complex provides experimental evidence that the pteridine ring of dihydrofolate is polarized when bound to the enzyme. We propose that this polarization of dihydrofolate acts as the driving force for protonation of the electron-rich O4 atom which occurs in the presence of NADPH. After this protonation of the substrate, a network of hydrogen bonds between O4, N5 and a bound water molecule facilitates transfer of the proton to N5 and transfer of a hydride ion from NADPH to the C6 atom to complete the reduction process.  相似文献   

20.
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   

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