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.
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   

2.
Achieving molecular recognition of NADPH binding sites is a compelling strategy to control many redox biological processes. The NADPH sites recognize the ubiquitous NADPH cofactor via highly conserved binding interactions, despite differences in the regulation of the hydride transfer in redox active proteins. We recently developed a photoactive NADPH substitute, called nanotrigger NT synchronizing the initiation of enzymatic catalysis of the endothelial NO‐synthase (eNOS) with a laser pulse. Spatial and temporal control of enzymatic activity by such a designed light‐driven activator would benefit from achieving molecular selectivity, i.e. activation of a single NADPH‐mediated enzyme. In this work, we probe the ability of NT to discriminate between two NADPH sites with light. The selected NADPH sites belong to dihydrofolate reductase dihydrofolate reductase enzyme (DHFR) and endothelial NO‐synthase (eNOS). Ultrafast kinetics showed that NT could not activate DHFR catalysis with a laser pulse in contrast with the observed trigger of eNOS catalysis leading to NO formation. Homology modelling, molecular dynamics simulations showed that NT discriminated between the two NADPH sites by different donor to acceptor distances and by local steric effects hindering light activation of DHFR catalysis. The data suggested that the narrow NADPH site required a tight fit of the nanotrigger at a suitable distance/angle to the electron acceptor for a specific activation of the catalysis. The ability of the nanotrigger to activate eNOS combined with a low reactivity in unfavourable NADPH sites makes NT a highly promising tool for targeting eNOS in endothelial cells with a laser pulse. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

3.
The 2.3-A crystal structure of recombinant human dihydrofolate reductase (EC 1.5.1.3, DHFR) has been solved as a binary complex with folate (a poor substrate at neutral pH) and also as a binary complex with an inhibitor, 5-deazafolate. The inhibitor appears to be protonated at N8 on binding, whereas folate is not. Rotation of the peptide plane joining I7 and V8 from its position in the folate complex permits hydrogen bonding of 5-deazafolate's protonated N8 to the backbone carbonyl of I7, thus contributing to the enzyme's greater affinity for 5-deazafolate than for folate. In this respect it is likely that bound 5-deazafolate furnishes a model for 7,8-dihydrofolate binding and, in addition, resembles the transition state for folate reduction. A hypothetical transition-state model for folate reduction, generated by superposition of the DHFR binary complexes human.5-deazafolate and chicken liver.NADPH, reveals a 1-A overlap of the binding sites for folate's pteridine ring and the dihydronicotinamide ring of NADPH. It is proposed that this binding-site overlap accelerates the reduction of both folate and 7,8-dihydrofolate by simultaneously binding substrate and cofactor with a sub van der Waals separation that is optimal for hydride transfer.  相似文献   

4.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

5.
Chopra S  Lynch R  Kim SH  Jackson M  Howell EE 《Biochemistry》2006,45(21):6596-6605
R67 dihydrofolate reductase (DHFR) is a novel homotetrameric protein that possesses 222 symmetry and a single, voluminous active site pore. This symmetry poses numerous limitations on catalysis; for example, two dihydrofolate (DHF) molecules or two NADPH molecules, or one substrate plus one cofactor can bind. Only the latter combination leads to catalysis. To garner additional information on how this enzyme facilitates transition-state formation, the temperature dependence of binding and catalysis was monitored. The binding of NADPH and DHF is enthalpy-driven. Previous primary isotope effect studies indicate hydride transfer is at least partially rate-determining. Accordingly, the activation energy associated with transition-state formation was measured and is found to be 6.9 kcal/mol (DeltaH(++)(25) = 6.3 kcal/mol). A large entropic component is also found associated with catalysis, TDeltaS(++)(25) = -11.3 kcal/mol. The poor substrate, dihydropteroate, binds more weakly than dihydrofolate (DeltaDeltaG = 1.4 kcal/mol) and displays a large loss in the binding enthalpy value (DeltaDeltaH = 3.8 kcal/mol). The k(cat) value for dihydropteroate reduction is decreased 1600-fold compared to DHF usage. This effect appears to derive mostly from the DeltaDeltaH difference in binding, demonstrating that the glutamate tail is important for catalysis. This result is surprising, as the para-aminobenzoyl-glutamate tail of DHF has been previously shown to be disordered by both NMR and crystallography studies. Viscosity studies were also performed and confirmed that the hydride transfer rate is not sensitive to sucrose addition. Surprisingly, binding of DHF, by both K(m) and K(d) determination, was found to be sensitive to added viscogens, suggesting a role for water in DHF binding.  相似文献   

6.
Polshakov VI  Birdsall B  Feeney J 《Biochemistry》1999,38(48):15962-15969
NMR measurements have been used to investigate rates of ring-flipping and the activation parameters for the trimethoxybenzyl ring of the antibacterial drug trimethoprim (TMP) bound to Lactobacillus casei dihydrofolate reductase (DHFR) for a series of ternary complexes formed with analogues of the coenzyme NADPH. Rates were obtained at several temperatures from line shape analyses ((13)C-edited HSQC (1)H spectra) and transfer of magnetization measurements (zz-HSQC) on complexes containing 3'-O-[(13)C]trimethoprim. Examination of the structures of the complexes indicates that ring-flipping can only be achieved following major conformational changes and transient fluctuations of the protein and coenzyme structure around the trimethoxybenzyl ring. There is no simple correlation between rates of ring-flipping and binding constants. The presence of the coenzyme nicotinamide ring (in either its reduced or its oxidized forms) in the binding site close to the trimethoxybenzyl ring moiety is the major factor reducing the ring-flipping on coenzyme binding. Thus, the ternary complex with NADPH shows the largest reduction in the rate of ring-flipping (11 +/- 3 s(-)(1) at 298 K) as compared with the binary complex (793 +/- 80 s(-)(1) at 298 K). Complexes with NADPH analogues that either have no nicotinamide ring or are known to have their nicotinamide rings removed from the binding site show the smallest reductions. For the DHFR.TMP.NADP(+) complex where there are two conformations present, very different rates of ring-flipping were observed for the two forms. The activation parameters (DeltaH() and DeltaS()) for the ring-flipping in all the complexes are discussed in terms of the protein-ligand interactions and the possible constraints on the pathway through the transition state.  相似文献   

7.
A kinetic scheme is presented for Lactobacillus casei dihydrofolate reductase that predicts steady-state kinetic parameters. This scheme was derived from measuring association and dissociation rate constants and pre-steady-state transients by using stopped-flow fluorescence and absorbance spectroscopy. Two major features of this kinetic scheme are the following: (i) product dissociation is the rate-limiting step for steady-state turnover at low pH and follows a specific, preferred pathway in which tetrahydrofolate (H4F) dissociation occurs after NADPH replaces NADP+ in the ternary complex; (ii) the rate constant for hydride transfer from NADPH to dihydrofolate (H2F) is rapid (khyd = 430 s-1), favorable (Keq = 290), and pH dependent (pKa = 6.0), reflecting ionization of a single group. Not only is this scheme identical in form with the Escherichia coli kinetic scheme [Fierke et al. (1987) Biochemistry 26, 4085] but moreover none of the rate constants vary by more than 40-fold despite there being less than 30% amino acid homology between the two enzymes. This similarity is consistent with their overall structural congruence. The role of Trp-21 of L. casei dihydrofolate reductase in binding and catalysis was probed by amino acid substitution. Trp-21, a strictly conserved residue near both the folate and coenzyme binding sites, was replaced by leucine. Two major effects of this substitution are on (i) the rate constant for hydride transfer which decreases 100-fold, becoming the rate-limiting step in steady-state turnover, and (ii) the affinities for NADPH and NADP+ which decrease by approximately 3.5 and approximately 0.5 kcal mol-1, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
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   

9.
Single photon counting pulse fluorimetry has been used in order to study the two ternary complexes GDH-GTP-NADPH and GDH-L-glutamate-NADPH and the quaternary complex GDH-GTP-L-glutamate-NADPH. The fluorescence decay of the enzyme-bound NADPH is not monoexponential in any of these complexes. Moreover, it does not seem to be dependent on the coenzyme concentration. The experimental curves can be satisfactorily fitted with the sum of two exponentials, the relative amplitudes of which significantly depend on the complex studied. Thus, for dihydronicotinamide two possible environments might exist in the enzyme active sites. It is also shown that the fluorescence decay times of the enzyme are shortened by the bound NADPH.  相似文献   

10.
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   

11.
This paper reviews the results from hybrid quantum/classical molecular dynamics simulations of the hydride transfer reaction catalysed by wild-type (WT) and mutant Escherichia coli and WT Bacillus subtilis dihydrofolate reductase (DHFR). Nuclear quantum effects such as zero point energy and hydrogen tunnelling are significant in these reactions and substantially decrease the free energy barrier. The donor-acceptor distance decreases to ca 2.7 A at transition-state configurations to enable the hydride transfer. A network of coupled motions representing conformational changes along the collective reaction coordinate facilitates the hydride transfer reaction by decreasing the donor-acceptor distance and providing a favourable geometric and electrostatic environment. Recent single-molecule experiments confirm that at least some of these thermally averaged equilibrium conformational changes occur on the millisecond time-scale of the hydride transfer. Distal mutations can lead to non-local structural changes and significantly impact the probability of sampling configurations conducive to the hydride transfer, thereby altering the free-energy barrier and the rate of hydride transfer. E. coli and B. subtilis DHFR enzymes, which have similar tertiary structures and hydride transfer rates with 44% sequence identity, exhibit both similarities and differences in the equilibrium motions and conformational changes correlated to hydride transfer, suggesting a balance of conservation and flexibility across species.  相似文献   

12.
The importance of salt bridge interactions at the NADPH binding site of dihydrofolate reductase has been studied by using site-directed mutagenesis. The mutations R44L and H45Q respectively disrupt the ionic contacts made between the 2'-phosphate and pyrophosphoryl moiety of the coenzyme and the N-terminal region of helix C. Equilibrium fluorescence experiments indicate that while the overall binding of NADPH to both free mutants is weakened by 1.1 and 1.5 kcal/mol (H45Q and R44L, respectively), the binding of dihydrofolate and tetrahydrofolate is unaffected. Despite the similar binding energies for both mutants, the transition state for the chemical hydride step is differentially destabilized relative to wild type (0.6 and 1.8 kcal/mol for H45Q and R44L, respectively). Both stopped-flow and pre-steady-state experiments suggest that the root of this effect may lie in multiple conformations for the E-NADPH complex of R44L. The ability of both mutants to transmit their effects beyond the local environment of the NADPH pocket is manifested in several details: (1) the pKa of Asp-27 (25 A away from the sites of mutation) is elevated from 6.5 in the wild type to 7.5 and 8.4 in H45Q and R44L, respectively; (2) NADPH elevates the off rates for tetrahydrofolate from 12 s-1 in the wild type to greater than 45 s-1 in R44L; and (3) bound tetrahydrofolate decreases the affinity of the enzymes for NADPH as reflected in the Km from 2 to 40 microM for H45Q (similar to wild type) but from 8 to 5000 microM for R44L.  相似文献   

13.
The variable residue Leu-28 of Escherichia coli dihydrofolate reductase (DHFR) and the corresponding residue Phe-31 in murine DHFR were interchanged, and the impact on catalysis was evaluated by steady-state and pre-steady-state analysis. The E. coli L28F mutant increased the pH-independent kcat from 11 to 50 s-1 but had little effect on Km(H2F). An increase in the rate constant for dissociation of H4F from E.H4F.NH (from 12 to 80 s-1) was found to be largely responsible for the increase in kcat. Unexpectedly, the rate constant for hydride transfer increased from 950 to 4000 s-1 with little perturbation of NADPH and NADP+ binding to E. Consequently, the flux efficiency of the E. coli L28F mutant rose from 15% to 48% and suggests a role in genetic selection for this variable side chain. The murine F31L mutant decreased the pH-independent kcat from 28 to 4.8 s-1 but had little effect on Km(H2F). A decrease in the rate constant for dissociation of H4F from E.H4F.NH (from 40 to 22 s-1) and E.H4F (from 15 to 0.4 s-1) was found to be mainly responsible for the decrease in kcat. The rate constant for hydride transfer decreased from 9000 to 5000 s-1 with minor perturbation of NADPH binding. Thus, the free energy differences along the kinetic pathway were generally similar in magnitude but opposite in direction to those incurred by the E. coli L28F mutant. This conclusion implies that DHFR hydrophobic active-site side chains impart their characteristics individually and not collectively.  相似文献   

14.
Rajagopalan PT  Lutz S  Benkovic SJ 《Biochemistry》2002,41(42):12618-12628
Recently, the participation in catalysis of residues spatially removed from the enzyme's active site has received considerable attention. The influence of the distal Gly-121 residue on the chemical step of hydride transfer in dihydrofolate reductase (DHFR) catalysis had been demonstrated previously [Cameron, C. E., and Benkovic, S. J. (1997) Biochemistry 36, 15792-15800]. In our continuing effort to identify catalytically important residues that are distal from the active site, we used sequence conservation information, kinetic data on site-directed mutants, dynamic motion information from NMR methods, and correlated motions from MD simulations to identify a subset of residues. Among them, the region spanning positions 41-45 is distal to the active site and was chosen as the focus for the mutagenesis and kinetic studies reported here. Specifically, the highly conserved Met-42 was selected for site-directed mutagenesis. While the reaction kinetics for the M42F mutant enzyme did not deviate from wild-type behavior, a 41-fold reduction in the forward hydride transfer rate was found for the M42W mutant. Given the established role of Gly-121 in the hydride transfer process, double mutant enzymes involving positions 42 and 121 were constructed and characterized. These double mutant enzymes generally showed little changes in substrate and cofactor binding but synergistic decreases in forward hydride transfer rates, while the decreases in reverse rates were additive. Along with supporting information from mixed quantum/classical MD simulations [Agarwal, P. K., Billeter, S. R., Rajagopalan, P. T., Benkovic, S. J., and Hammes-Schiffer, S. (2002) Proc. Natl. Acad. Sci. U.S.A. 99, 2794-2799], the data suggest that a coupled dynamic motion of these distal residues enhances crossing of the chemical reaction barrier and imply an expanded nonstatic role for the protein fold in catalysis.  相似文献   

15.
J Thillet  J A Adams  S J Benkovic 《Biochemistry》1990,29(21):5195-5202
A kinetic mechanism is presented for mouse dihydrofolate reductase that predicts all the steady-state parameters and full time-course kinetics. This mechanism was derived from association and dissociation rate constants and pre-steady-state transients by using stopped-flow fluorescence and absorbance measurements. The major features of this kinetic mechanism are as follows: (1) the two native enzyme conformers, E1 and E2, bind ligands with varying affinities although only one conformer, E1, can support catalysis in the forward direction, (2) tetrahydrofolate dissociation is the rate-limiting step under steady-state turnover at low pH, and (3) the pH-independent rate of hydride transfer from NADPH to dihydrofolate is fast (khyd = 9000 s-1) and favorable (Keq = 100). The overall mechanism is similar in form to the Escherichia coli kinetic scheme (Fierke et al., 1987), although several differences are observed: (1) substrates and products predominantly bind the same form of the E. coli enzyme, and (2) the hydride transfer rate from NADPH to either folate or dihydrofolate is considerably faster for the mouse enzyme. The role of Glu-30 (Asp-27 in E. coli) in mouse DHFR has also been examined by using site-directed mutagenesis as a potential source of these differences. While aspartic acid is strictly conserved in all bacterial DHFRs, glutamic acid is conserved in all known eucaryotes. The two major effects of substituting Asp for Glu-30 in the mouse enzyme are (1) a decreased rate of folate reduction and (2) an increased rate of hydride transfer from NADPH to dihydrofolate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
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   

17.
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   

18.
This brief review discusses our current understanding of the molecular basis of enzyme catalysis. A historical development is presented, beginning with steady state kinetics and progressing through modern fast reaction methods, nuclear magnetic resonance, and single-molecule fluorescence techniques. Experimental results are summarized for ribonuclease, aspartate aminotransferase, and especially dihydrofolate reductase (DHFR). Multiple intermediates, multiple conformations, and cooperative conformational changes are shown to be an essential part of virtually all enzyme mechanisms. In the case of DHFR, theoretical investigations have provided detailed information about the movement of atoms within the enzyme-substrate complex as the reaction proceeds along the collective reaction coordinate for hydride transfer. A general mechanism is presented for enzyme catalysis that includes multiple intermediates and a complex, multidimensional standard free energy surface. Protein flexibility, diverse protein conformations, and cooperative conformational changes are important features of this model.  相似文献   

19.
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   

20.
A fluorescently-labeled, conformationally-sensitive Bacillus stearothermophilus (Bs) dihydrofolate reductase (DHFR) (C73A/S131CMDCC DHFR) was developed and used to investigate kinetics and protein conformational motions associated with methotrexate (MTX) binding. This construct bears a covalently-attached fluorophore, N-[2-(1-maleimidyl)ethyl]-7-(diethylamino)coumarin-3-carboxamide (MDCC) attached at a distal cysteine, introduced by mutagenesis. The probe is sensitive to the local molecular environment, reporting on changes in the protein structure associated with ligand binding. Intrinsic tryptophan fluorescence of the unlabeled Bs DHFR construct (C73A/S131C DHFR) also showed changes upon MTX association. Stopped-flow analysis of all data can be understood by invoking the presence of two native state DHFR conformers that bind to MTX at different rates (20.2 and 0.067 μM−1 s−1), similar to previously published findings for Escherichia coli DHFR. Probe fluorescence of C73A/S131CMDCC DHFR predominantly reports on MTX binding to one of the conformers while intrinsic tryptophan fluorescence of C73A/S131C DHFR reports on binding to the other conformer. This study demonstrates the use of an extrinsic fluorophore attached to a distal region to investigate ligand binding interactions that are not experimentally accessible via intrinsic tryptophan fluorescence alone. The thermostability of C73A/S131CMDCC DHFR provides an important new tool with applications for investigating the temperature dependence of DHFR conformational changes associated with binding and catalysis.  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