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.
A remarkable correlation has been discovered between fluorescence lifetimes of bound NADPH and rates of hydride transfer among mutants of dihydrofolate reductase (DHFR) from Escherichia coli. Rates of hydride transfer from NADPH to dihydrofolate change by a factor of 1,000 for the series of mutant enzymes. Since binding constants for the initial complex between coenzyme and DHFR change by only a factor of 10, the major portion of the change in hydride transfer must be attributed to losses in transition-state stabilization. The time course of fluorescence decay for NADPH bound to DHFR is biphasic. Lifetimes ranging from 0.3 to 0.5 ns are attributed to a solvent-exposed dihydronicotinamide conformation of bound coenzyme which is presumably not active in catalysis, while decay times (tau 2) in the range of 1.3 to 2.3 ns are assigned to a more tightly bound species of NADPH in which dihydronicotinamide is sequestered from solvent. It is this slower component that is of interest. Ternary complexes with three different inhibitors, methotrexate, 5-deazafolate, and trimethoprim, were investigated, along with the holoenzyme complex; 3-acetylNADPH was also investigated. Fluorescence polarization decay, excitation polarization spectra, the temperature variation of fluorescence lifetimes, fluorescence amplitudes, and wavelength of absorbance maxima were measured. We suggest that dynamic quenching or internal conversion promotes decay of the excited state in NADPH-DHFR. When rates of hydride transfer are plotted against the fluorescence lifetime (tau 2) of tightly bound NADPH, an unusual correlation is observed. The fluorescence lifetime becomes longer as the rate of catalysis decreases for most mutants studied. However, the fluorescence lifetime is unchanged for those mutations that principally alter the binding of dihydrofolate while leaving most dihydronicotinamide interactions relatively undisturbed. The data are interpreted in terms of possible dynamic motions of a flexible loop region in DHFR which closes over both substrate and coenzyme binding sites. These motions could lead to faster rates of fluorescence decay in holoenzyme complexes and, when correlated over time, may be involved in other motions which give rise to enhanced rates of catalysis in DHFR.  相似文献   

2.
The interaction of type II R67 dihydrofolate reductase (DHFR) with its cofactor nicotinamide adenine dinucleotide phosphate (NADP(+)) has been studied using nuclear magnetic resonance (NMR). Doubly labeled [U-(13)C,(15)N]DHFR was obtained from Escherichia coli grown on a medium containing [U-(13)C]-D-glucose and (15)NH(4)Cl, and the 16 disordered N-terminal amino acids were removed by treatment with chymotrypsin. Backbone and side chain NMR assignments were made using triple-resonance experiments. The degeneracy of the amide (1)H and (15)N shifts of the tetrameric DHFR was preserved upon addition of NADP(+), consistent with kinetic averaging among equivalent binding sites. Analysis of the more titration-sensitive DHFR amide resonances as a function of added NADP(+) gave a K(D) of 131 +/- 50 microM, consistent with previous determinations using other methodology. We have found that the (1)H spectrum of NADP(+) in the presence of the R67 DHFR changes as a function of time. Comparison with standard samples and mass spectrometric analysis indicates a slow conversion of NADP(+) to NAD(+), i.e., an apparent NADP(+) phosphatase activity. Studies of this activity in the presence of folate and a folate analogue support the conclusion that this activity results from an interaction with the DHFR rather than a contaminating phosphatase. (1)H NMR studies of a mixture of NADP(+) and NADPH in the presence of the enzyme reveal that a ternary complex forms in which the N-4A and N-4B nuclei of the NADPH are in the proximity of the N-4 and N-5 nuclei of NADP(+). Studies using the NADP(+) analogue acetylpyridine adenosine dinucleotide phosphate (APADP(+)) demonstrated a low level of enzyme-catalyzed hydride transfer from NADPH. Analysis of DHFR backbone dynamics revealed little change upon binding of NADP(+). These additional catalytic activities and dynamic behavior are in marked contrast to those of type I DHFR.  相似文献   

3.
Fluorescent emission kinetics of isolated spinach chloroplasts have been observed at room temperature with an instrument resolution time of 10 ps using a frequency doubled, mode-locked Nd:glass laser and an optical Kerr gate. At 685 nm two maxima are apparent in the time dependency of the fluorescence; the first occurs at 15 ps and the second at 90 ps after the flash. The intervening minimum occurs at about 50 ps. On the basis of theoretical models, lifetimes of the components associated with the two peaks and spectra (in escarole chloroplasts), the fluorescence associated with the first peak is interpreted as originating from Photosystem I (PSI) (risetime ≤10 ps, lifetime ≤10 ps) and the second peak from Photosystem II (PSII) (lifetime, 210 ps in spinach chloroplasts and 320 ps in escarole chloroplasts). The fact that there are two fluorescing components with a quantum yield ratio ≤0.048 explains the previous discrepancy between the quantum yield of fluorescence measured in chloroplasts directly and that calculated from the lifetime of PSII. The 90 ps delay in the peak of PSII fluorescence is probably explained by energy transfer between accessory pigments such as carotenoids and Chl a. Energy spillover between PSI and PSII is not apparent during the time of observation. The results of this work support the view that the transfer of excitation energy to the trap complex in both photosystems occurs by means of a molecular excitation mechanism of intermediate coupling strength. Although triplet states are not of major importance in energy transfer to PSII traps, the possibility that they are involved in PSI photochemistry has not been eliminated.  相似文献   

4.
Fluorescence spectrum of camel lens zeta-crystallin, a major protein in the lens of camelids and histicomorph rodents, showed maximum emission at 315 nm. This emission maximum is blue shifted compared to most proteins, including alpha-crystallin, and appeared to be due to tryptophan in highly hydrophobic environment. Interaction of NADPH with zeta-crystallin quenched the protein fluorescence and enhanced the fluorescence of bound NADPH. Analysis of fluorescence quenching suggested high-affinity interaction between NADPH and zeta-crystallin with an apparent Km<0.45 microM. This value is at least an order of magnitude lower than that suggested by activity measurements. Analysis of NADPH fluorescence showed a biphasic curve representing fluorescence of free- and bound-NADPH. The intersection between free- and bound-NADPH closely paralleled the enzyme concentration, suggesting one mole of NADPH was bound per subunit of the enzyme. Phenanthrenequinone (PQ), the substrate of zeta-crystallin, also was able to quench the fluorescence of zeta-crystallin, albeit weaker than NADPH. Quantitative analysis suggested that zeta-crystallin had low affinity for PQ in the absence of NADPH, and PQ binding induced significant conformational changes in zeta-crystallin.  相似文献   

5.
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   

6.
The structure of mouse L1210 dihydrofolate reductase (DHFR) complexed with NADPH and trimethoprim has been refined at 2.0 A resolution. The analogous complex with NADPH and methotrexate has been refined at 2.5 A resolution. These structures reveal for the first time details of drug interactions with a mammalian DHFR, which are compared with those observed from previous X-ray investigations of DHFR/inhibitor complexes. The refined L1210 structure has been used as the basis for the construction of a model of the human enzyme. There are only twenty-one sequence differences between mouse L1210 and human DHFRs, and all but two of these are located close to the molecular surface: a strong indication that the active sites are essentially identical in these two mammalian enzymes.  相似文献   

7.
The apoenzyme of wild-type (WT) dihydrofolate reductase (DHRF) from Escherichia coli exists in two conformational states, Et and Ew, which differ in affinity for NADPH and in kinetic competence. Dissociation constants for the binary complex of NADPH with the two conformers differ by over 100-fold (KDt = 0.17 microM, KDw = 22 microM). Rate constants governing the interconversion of conformers are small (t1/2 for Ew----Et = 71 s), and since Ew is not catalytically competent, this conversion is accompanied by an increase in catalytic velocity. The equilibrium proportion of Et in the absence of ligands is 63%, but binding of NADPH greatly increases this proportion, and t1/2 for conversion of Ew.NADPH to Et.NADPH is 30 s. This conformational equilibrium has also been examined in mutant enzyme in which aspartate 27 is replaced by asparagine (D27N E. coli DHFR). Although ASp27 is an active site residue, it does not interact directly with bound NADPH, and in the mutant the rate constant for NADPH binding to Et is unchanged as are the dissociation constants for NADPH complexes with Et or Ew. However, for mutant apoenzyme, the proportion of Et is decreased to 18% in the absence of ligands so that the overall KD for NADPH is increased (0.15 microM for WT E. coli DHFR, 0.68 microM for D27N E. coli DHFR). The lower proportion of Et is due to a decreased rate for Ew----Et (t1/2 = 221 s) and an increased rate for Et----Ew (t1/2 = 50 s versus 120 s for WT E. coli DHFR).  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.
Pulse radiolytically generated halogenated alkylperoxyl radicals, namely CCl3OO*, CBr3OO*, CHCl2OO*, CHBr2OO*, etc. have been employed for a study of the oxidation of beta-carotene in a quaternary microemulsion. All these halogenated alkylperoxyl radicals produce a radical cation (absorption maximum at 840 nm), either via the initial formation of an adduct (absorption maximum at 740 nm), or by direct reaction. There was a considerable blue shift of both absorption peaks in the present system as compared to those earlier reported in non-polar as well as in micellar media. An additional intense absorption peak at approximately 345 nm was noted at a longer time scale and was stable up to 5 ms. On irradiation with a large number of electron pulses, a stable product with an absorption maximum at 315 nm appeared. On excitation at 315 nm, this product gave a fluorescence spectrum with an emission maximum at 525 nm. The absorption and fluorescence spectra of the stable product compared with those of retinol; this was further confirmed by HPLC analysis. A suitable mechanism has been proposed.  相似文献   

10.
Heyes DJ  Ruban AV  Hunter CN 《Biochemistry》2003,42(2):523-528
The light-driven enzyme NADPH:protochlorophyllide oxidoreductase (POR) catalyzes the reduction of protochlorophyllide (Pchlide) to chlorophyllide (Chlide), a key regulatory step in the chlorophyll biosynthesis pathway. As POR is light activated, it allows intermediates in the reaction pathway to be observed by initiating catalysis with illumination at low temperatures, a technique that has recently been used to study the initial photochemistry. Here, we use low-temperature spectroscopy to show that the catalytic mechanism of POR involves two additional steps, which do not require light and have been termed the "dark" reactions. The first of these involves the conversion of the product of the initial light-driven reaction, a nonfluorescent radical species, into a new intermediate that has an absorbance maximum at 681 nm and a fluorescence peak at 684 nm. During the second dark step this species gradually blue shifts to yield the product, Chlide. The temperature dependence for each of these two processes was measured; the data revealed that these steps could only occur close to or above the "glass transition" temperature of proteins, suggesting that domain movements and/or reorganization of the protein are required for these stages of the catalytic mechanism.  相似文献   

11.
Despite much work, many key aspects of the mechanism of the dihydrofolate reductase (DHFR) catalyzed reduction of dihydrofolate remain unresolved. In bacterial forms of DHFR both substrate and water access to the active site are controlled by the conformation of the mobile M20 loop. In vertebrate DHFRs only one conformation of the residues corresponding to the M20 loop has been observed. Access to the active site was proposed to be controlled by residue 31. MD simulations of chicken DHFR complexed with substrates and cofactor revealed a closing of the side chain of Tyr 31 over the active site on binding of dihydrofolate. This conformational change was dependent on the presence of glutamate on the para-aminobenzoylamide moiety of dihydrofolate. In its absence, the conformation remained open. Although water could enter the active site and hydrogen bond to N5 of dihydrofolate, indicating the feasibility of water as the proton donor, this was not controlled by the conformation of Tyr 31. The water accessibility of the active site was low for both conformations of Tyr 31. However, when hydride was transferred from NADPH to C6 of dihydrofolate before protonation, the average time during which water was found in hydrogen bonding distance to N5 of dihydrofolate in the active site increased almost fivefold. These results indicated that water can serve as the Broensted acid for the protonation of N5 of dihydrofolate during the DHFR catalyzed reduction.  相似文献   

12.
Fluorescence spectrum of camel lens ζ-crystallin, a major protein in the lens of camelids and histicomorph rodents, showed maximum emission at 315 nm. This emission maximum is blue shifted compared to most proteins, including α-crystallin, and appeared to be due to tryptophan in highly hydrophobic environment. Interaction of NADPH with ζ-crystallin quenched the protein fluorescence and enhanced the fluorescence of bound NADPH. Analysis of fluorescence quenching suggested high-affinity interaction between NADPH and ζ-crystallin with an apparent Km<0.45 μM. This value is at least an order of magnitude lower than that suggested by activity measurements. Analysis of NADPH fluorescence showed a biphasic curve representing fluorescence of free- and bound-NADPH. The intersection between free- and bound-NADPH closely paralleled the enzyme concentration, suggesting one mole of NADPH was bound per subunit of the enzyme. Phenanthrenequinone (PQ), the substrate of ζ-crystallin, also was able to quench the fluorescence of ζ-crystallin, albeit weaker than NADPH. Quantitative analysis suggested that ζ-crystallin had low affinity for PQ in the absence of NADPH, and PQ binding induced significant conformational changes in ζ-crystallin.  相似文献   

13.
Aromatase has been purified to homogeneity from human placental microsomes based on detection of its catalytic activities in the eluates from columns of octylamino-Sepharose 4B, hydroxylapatite, Mono S, hydroxylapatite HCA, and Mono Q. The purified preparation shows only one band corresponding to the apparent subunit molecular weight of 51,000 daltons on sodium dodecyl sulfate-polyacrylamide gel. The aromatase in the presence of NADPH and NADPH-cytochrome P-450 reductase converts testosterone to 17 beta-estradiol with the high specific activity of 103 nmol/min/mg of protein. However, whether the preparation is reduced by sodium dithionate chemically or by NADPH and the reductase enzymatically, its reduced, CO-difference spectrum has no peak at about 450 nm and has only a small peak at about 420 nm, probably due to its inactivation in spite of the catalytically full activity in the same preparation. The absolute spectrum of the aromatase exhibits a Soret peak at 423 nm in the absence of testosterone and addition of testosterone to the aromatase sample makes its absorption peak shift gradually from 423 to 393 nm (high spin type peak), which is a usual characteristic in the spectrum of cytochrome P-450. The reconstituted aromatase system efficiently catalyzes aromatization of 4-androstenedione, 19-hydroxy-4-androstenedione as well as testosterone. 16 alpha-Hydroxy-4-androstenedione and 16 alpha-hydroxytestosterone are also aromatized less efficiently and 19-nortestosterone is aromatized least efficiently. The reconstituted aromatase could scarcely oxidize various xenobiotics examined, suggesting a strict and narrow substrate specificity of this enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
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   

15.
Dihydrofolate reductase (DHFR) (5,6,7,8-THF: NADDP+ oxidoreductase, EC 1.5.1.3) was purified 205-fold to apparent homogeneity from the crude extracts of Lactobacillus leichmannii. It has UV absorption maxima at 280 nm, M(r) of 20,000, Stokes radius of 0.34 nm and a S20.w value of 0.12 S. The preparation showed the presence of 168 amino acid residues with threonine and lysine as the NH2- and COOH- terminal end-groups respectively and a single reactive sulfhydryl group. pCMB inhibited the enzyme activity (IC50 = 2 microM). The enzyme has a pH optimum of 7.4 and is thermally inactivated at > 35 degrees C. It is activated by 0.1 M KCl and KI and 2 M urea. 3-4 M urea completely inactivated the enzyme. Enzyme has Km values of 3.5 microM and 6.2 microM for NADPH and DHF respectively, and a Ki value of 7 nM for MTX, the inhibition being competitive.  相似文献   

16.
In Arabidopsis thaliana Por C has been identified only on sequence homology to that of por A and por B. To demonstrate its catalytic function Arabidopsis thaliana protochlorophyllide oxidoreductase C gene (por c) that codes for the mature part of POR C protein having 335 amino acids was expressed in Escherchia coli cells. The POR C enzyme in the presence of NADPH and protochlorophyllide when incubated in dark formed a ternary complex. When it was excited at 433 nm, it had a fluorescence emission peak at 636 nm. After illumination with actinic cool white fluorescent light, a peak at 673 nm due to chlorophyllide gradually increased with concomitant decrease of 636 nm emission, demonstrating the gradual phototransformation of protochlorophyllide to chlorophyllide. The significance of differential por gene expression in light and dark among different species is discussed.  相似文献   

17.
Isothermal titration calorimetry (ITC) was used to determine the thermodynamic driving force for inhibitor binding to the enzyme dihydrofolate reductase (DHFR) from Escherichia coli. 1,4-Bis-{[N-(1-imino-1-guanidino-methyl)]sulfanylmethyl}-3,6-dimethyl-benzene (1) binds DHFR:NADPH with a K(d) of 13±5 nM while the related inhibitor 1-{[N-(1-imino-guanidino-methyl)]sulfanylmethyl}-3-trifluoromethyl-benzene (2) binds DHFR:NADPH with a K(d) of 3.2±2.2 μM. The binding of these inhibitors has both a favorable entropy and enthalpy of binding. Additionally, we observe positive binding cooperativity between both 1 and 2 and the cofactor NADPH. Binding of compound 1 to DHFR is 285-fold tighter in the presence of the NADPH cofactor than in its absence. We did not detect binding of 2 to DHFR in the absence of NADPH. The backbone amide (1)H and (15)N NMR resonances of DHFR:NADPH and both DHFR:NADPH inhibitor complexes were assigned in order to better understand the binding of these inhibitors in solution. The chemical shift perturbations observed with the binding of 1 were greatest at residues closest to the binding site, but significant perturbations also occur away from the inhibitor location at amino acids in the vicinity of residue 58 and in the GH loop. The pattern of chemical shift changes observed with the binding of 2 is similar to that seen with 1. The main differences in chemical shift perturbation between the two inhibitors are in the Met20 loop and in residues at the interface between the inhibitor and NADPH.  相似文献   

18.
  • 1.1. Covalent coupling of fluorescein to methotrexate (MTX) by a 5-carbon spacer yields a dihydrofolate reductase (DHFR) inhibitor (FMTX) with Ki = 11 nM.
  • 2.2. FMTX shows a fluorescence quenching with respect to fluorescein which is relieved by binding to the enzyme.
  • 3.3. The dissociation constants (Kd) of MTX, FMTX, NADPH and 7,8-dihydrofolate (DHF) from bovine liver DHFR have been determined by fluorometric titrations.
  • 4.4. The Kd values for NADPH, MTX and FMTX from the complementary binary complexes (MTX·DHFR, FMTX·DHFR and NADPH·DHFR) were also obtained; these show a 2- to 4-fold decrease with respect to those obtained by titration of the free enzyme.
  • 5.5. A competitive assay for MTX has been developed by exploiting the fluorescence enhancement of DHFR-bound FMTX. This assay may be useful for the routine determination of MTX in the concentration range from 10−9 to 10−7 M.
  相似文献   

19.
Synthesis and biological evaluation of a fluorescent analogue of folic acid   总被引:4,自引:0,他引:4  
A fluorescein derivative of the lysine analogue of folic acid, N alpha-pteroyl-N epilson-(4'-fluoresceinthiocarbamoyl)-L-lysine (PLF), was synthesized as a probe for dihydrofolate reductase (DHFR) and a membrane folate binding protein (m-FBP). Excitation of PLF at 282 nm and at 497 nm gave a fluorescence emission maximum at 518 nm. Binding of PLF to human DHFR or human placental m-FBP results in approximately a 20-fold enhancement in the magnitude of the fluorescence emission, suggesting that the ligand interacts with a hydrophobic region on these proteins. Additional evidence suggests that an energy transfer may occur between the pteridine and the fluorescein moieties. PLF binds to the active site of human DHFR since methotrexate (MTX) competes stoichiometrically and the denatured enzyme in the presence of PLF did not exhibit fluorescent enhancement. The dissociation constant for the fluorescein derivative with respect to human DHFR is 115 nM as compared to 111 nM for folic acid. The Ki value for the competitive inhibition of human DHFR by the fluorescent analogue of folic acid is 2.0 microM compared to 0.48 microM for folic acid. PLF was reduced to N alpha-(7,8-dihydropteroyl)-N epilson-(4'-fluoresceinthiocarbamoyl)-L-lysine (H2PLF) and assayed by the enzymatic conversion to the tetrahydro derivative. The Km value for human DHFR for the dihydrofolate analogue is 2.0 microM. The KD value for H2PLF to human DHFR is 47 nM as compared to 44 nM for dihydrofolate. The KD values for both H2PLF and PLF indicate that the fluorescein moiety does not significantly affect folate binding in enzyme binary complexes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Smiley RD  Stinnett LG  Saxton AM  Howell EE 《Biochemistry》2002,41(52):15664-15675
R67 dihydrofolate reductase (DHFR) is an enzyme, encoded by an R-plasmid, that confers resistance to the antibacterial agent trimethoprim. This homotetramer possesses a single active site pore and exact 222 symmetry. The symmetry imposes constraints on the ability of the enzyme to optimize binding of the substrate, dihydrofolate (DHF), and the cofactor, NADPH, resulting in a "one site fits both ligands" approach. This approach allows formation of either a NADPH.NADPH, dihydrofolate.dihydrofolate, or NADPH.dihydrofolate complex. The first two complexes are nonproductive, while the third is the productive catalytic species. To break the symmetry of the active site, a tandem array of four R67 DHFR genes has been linked in frame, allowing individual manipulation of each gene copy. Various numbers and combinations of asymmetric Q67H mutations have been engineered into the tandem gene array. The Q67H mutation was chosen for investigation as it was previously found to tighten binding to both dihydrofolate and NADPH by approximately 100-fold in homotetrameric R67 DHFR [Park, H., Bradrick, T. D., and Howell, E. E. (1997) Protein Eng. 10, 1415-1424]. Nonadditive effects on ligand binding are observed when one to four mutations are inserted, indicating either conformational changes in the protein or different cooperativity patterns in the ligand-ligand interactions. From steady state kinetics, addition of Q67H mutations does not drastically affect formation of the NADPH.dihydrofolate complex; however, a large energy difference between the productive and nonproductive complexes is no longer maintained. A role for Q67 in discriminating between these various states is proposed. Since theories of protein evolution suggest gene duplication followed by accumulation of mutations can lead to divergence of activity, this study is a first step toward asking if introduction of asymmetric mutations in the quadrupled R67 DHFR gene can lead to optimization of ligand binding sites.  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