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 546 毫秒
1.
Interaction of several representative folate, quinazoline and pyridine nucleotide derivatives with dihydrofolate reductase from amethopterin-resistant Lactobacillus casei induces dramatic changes in its circular dichroic spectral properties. The binding of dihydrofolate induces a large extrinsic Cotton effect at 295 nm ([theta] = 113 800 deg . cm2 . dm-1). The generation of this band by dihydrofolate is strictly dependent on complex formation with a single substrate binding site and a KD = 7 . 10(-6) M. The other binary complexes examined include the enzyme . NADPH, enzyme . amethopterin, enzyme . folate, and enzyme . methasquin. All such complexes differ in spectral detail, the negative ellipticity at 330 nm being characteristic of the "folate site" complexes. The circular dichroic spectrum of the ternary complex of reductase . NADPH . methotrexate shows a positive symmetrical band centered at 360 nm ([theta] - 32 000 deg . cm2 . dm-1). Since both of the corresponding binary complexes exhibit negative bands in this region, this induced band represents a unique molecular property of the ternary complex. Chemical modification of a single tryptophan residue of the enzyme, as determined from magnetic circular dichroism spectra, results in a complete loss in the ability to bind either dihydrofolate or NADPH.  相似文献   

2.
The thermodynamic parameters of the binding of antifolate drugs to bovine liver dihydrofolate reductase (EC 1.5.1.3., 5,6,7,8-tetrahydrofolate: NADP+ oxidoreductase) have been measured with a flow microcalorimetric method. These parameters are greatly influenced by the structure of the inhibitor and/or by the presence of NADPH and above all by temperature. For all the compounds studied, binding at 37 degrees C is driven by favourable enthalpy variations, whereas entropy variations are unfavourable. At 10 degrees C, reactions are both enthalpically and entropically driven. These effects can be explained by a partial thermal denaturation of dihydrofolate reductase at 37 degrees C, which is restructured by NADPH and/or the antifolate. The refolding induced by the antifolate trimetrexate may explain its high association constant in the binary system (without NADPH), and the weaker cooperative effect of NADPH in the ternary system, as compared to methotrexate. In contrast, the poor affinity of trimethoprim for mammalian dihydrofolate reductase in binary and ternary systems at 37 degrees C is the result of a weaker stabilizing effect of this compound as regards temperature increase. Heat capacity variation linked to the complex formation reaction showed that this conformational transition is more pronounced between 25 and 37 degrees C than between 10 and 25 degrees C. Thus, the ability of the inhibitors to give to dihydrofolate reductase a more stable thermal behaviour at 37 degrees C is determinant in their binding.  相似文献   

3.
Dihydrofolate reductase (EC 1.5.1.3), purified to homogeneity from an amethopterin-resistant subline (R6) of cultured L1210 murine leukemia cells, has been used to study enzyme-substrate and enzyme-inhibitor complexes. NADPH, NADP+acid-modified NADPH (λmax at 265 nm, elevated absorbance at 290 nm), 2′-phosphoadenosine-5′-diphosphate ribose, dihydrofolate, and amethopterin formed binary complexes with the enzyme. Ternary complexes could be formed by admixing the enzyme with: (a) NADPH and amethopterin; (b) NADP+ and tetahydrofolate; and (c) acid-modified NADPH and dihydrofolate. All of these complexes migrated as stable well-defined bands on polyacrylamide gel electrophoresis at pH 8.3. The bands could be visualized by staining both for enzyme activity and for protein. These binary and ternary complexes were also stable to extensive dialysis. Spectra of the dialyzed enzyme complexes indicated that each ligand was present at an equimolar ratio with the enzyme.  相似文献   

4.
The fluorescence emission spectrum of soybean dihydrofolate reductase suggests that the emitting tryptophan residues are situated in a hydrophobic microenvironment. The dissociation constants determined from fluorescence and circular dichroism data reveal that the soybean enzyme has a lower affinity for substrates and substrate analogs than that determined for dihydrofolate reductases isolated from other sources. The binding of methotrexate to the soybean enzyme does not affect the binding of NADPH. Similarly, the binding of NADPH has no effect on subsequent methotrexate binding. Polarimetric study indicates that the enzyme has a low (ca. 5%) α-helical content. Addition of dihydrofolate to the soybean enzyme results in the generation of a positive ellipticity band at 298 nm with a molar ellipticity, [θ], of 186,000, whereas the binding of folate induces a negative ellipticity band at 280 nm with [θ] of ?181,000. The qualitative and quantitative differences in the circular dichroism of the enzyme-dihydrofolate and enzyme-folate complexes indicate that the mode of binding of these ligands may be different. The formation of an enzyme-NADPH complex is accompanied by a negative Cotton effect at 270 nm. These studies indicate that the binding of substrates or inhibitors causes significant conformational changes in the enzyme and also leads to the formation of a number of spectroscopically identifiable complexes.  相似文献   

5.
13C nuclear magnetic resonance spectra have been obtained for complexes of [2-13C]methotrexate and [2-13C]trimethoprim with wild-type dihydrofolate reductase (DHFR) from Escherichia coli and with two mutant enzymes in which aspartic acid-27 is replaced by asparagine and by serine, respectively. In both the wild-type and mutated enzymes, exchange between the free inhibitor and the enzyme-complexed inhibitor is slow on the NMR time scale; hence, despite the considerably increased dissociation constants for binary complexes with the enzymes, the dissociation rate remains small relative to the frequency separation of the resonances. In all cases but one, the pKa of an inhibitor that is complexed to enzyme differs greatly from that of the free inhibitor. However, while the pKa of both inhibitors in complexes with the wild-type enzyme is elevated to above 10, the pKa of the inhibitors complexed with the Asn-27 and Ser-27 enzymes is lowered to a value below 4. Exact determinations of bound pKa values are limited by the solubility of the enzyme and the dissociation constants of the complexes. The single exception to these general conclusions is the ternary complex of the Ser-27 DHFR with trimethoprim and NADPH. In this complex, both free and enzyme-complexed trimethoprim exhibit similar pKa values (approximately equal to 7.6). However, both the exchange between free and enzyme-complexed inhibitor and the protonation of the enzyme-complexed inhibitor are slow in the NMR time scale, so that the spectra reveal three resonances corresponding to free inhibitor, to protonated enzyme-complexed inhibitor, and to unprotonated enzyme-complexed inhibitor.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
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   

7.
Circular dichroism has been used to monitor the binding of pyridine nucleotide cofactors to enzyme-folate analog complexes of dihydrofolate reductase from Escherichia coli B (MB 1428). The enzyme binds one molar equivalent of many folate analogs and two molar equivalents of several pyridine nucleotide cofactors. The apo-enzyme has very low optical activity. The binding of folate analogs including folate, dihydrofolate, methotrexate, trimethoprim and pyrimethamine induce large Cotton effects. Pyridine nucleotides when bound to the enzyme-folate analog complexes also induce new optically active bands; all the effects being due to the first molar equivalent of cofactor bound. NADPH and NADP+ induce very similar bands when bound to the enzyme-methotrexate complex suggesting that the geometry of the complexes formed are very similar. The oxidized and reduced cofactor likewise have similar effects on the enzyme-folate complex. However, NADPH and NADP+ addition to both the enzyme-trimethoprim and enzyme-pyrimethamine complexes have significantly different effects on the circular dichroism spectra, suggesting that the inhibitors which are less homologous to the natural dihydrofolate substrate allow more conformational freedom in the enzyme-inhibitor-cofactor complex. In most cases the prior binding of the folate analog greatly increases the binding of the first molar equivalent of cofactor so that at concentrations of approx. 5-20 muM the binding appears stoichiometric. Pyrimethamine is an exception in that it apparently has no effect on the binding of NADPH to the enzyme.  相似文献   

8.
Pteridine reductase (PTR1) is an NADPH-dependent short-chain reductase found in parasitic trypanosomatid protozoans. The enzyme participates in the salvage of pterins and represents a target for the development of improved therapies for infections caused by these parasites. A series of crystallographic analyses of Leishmania major PTR1 are reported. Structures of the enzyme in a binary complex with the cofactor NADPH, and ternary complexes with cofactor and biopterin, 5,6-dihydrobiopterin, and 5,6,7,8-tetrahydrobiopterin reveal that PTR1 does not undergo any major conformational changes to accomplish binding and processing of substrates, and confirm that these molecules bind in a single orientation at the catalytic center suitable for two distinct reductions. Ternary complexes with cofactor and CB3717 and trimethoprim (TOP), potent inhibitors of thymidylate synthase and dihydrofolate reductase, respectively, have been characterized. The structure with CB3717 reveals that the quinazoline moiety binds in similar fashion to the pterin substrates/products and dominates interactions with the enzyme. In the complex with TOP, steric restrictions enforced on the trimethoxyphenyl substituent prevent the 2,4-diaminopyrimidine moiety from adopting the pterin mode of binding observed in dihydrofolate reductase, and explain the inhibition properties of a range of pyrimidine derivates. The molecular detail provided by these complex structures identifies the important interactions necessary to assist the structure-based development of novel enzyme inhibitors of potential therapeutic value.  相似文献   

9.
The chemical shifts of all the aromatic proton and anomeric proton resonances of NADP+, NADPH, and several structural analogues have been determined in their complexes with Lactobacillus casei dihydrofolate reductase by double-resonance (saturation transfer) experiments. The binding of NADP+ to the enzyme leads to large (0.9-1.6 ppm) downfield shifts of all the nicotinamide proton resonances and somewhat smaller upfield shifts of the adenine proton resonance. The latter signals show very similar chemical shifts in the binary and ternary complexes of NADP+ and the binary complexes of several other coenzymes, suggesting that the environment of the adenine ring is similar in all cases. In contrast, the nicotinamide proton resonances show much greater variability in position from one complex to another. The data show that the environments of the nicotinamide rings of NADP+, NADPH, and the thionicotinamide and acetylpyridine analogues of NADP+ in their binary complexes with the enzyme are quite markedly different from one another. Addition of folate or methotrexate to the binary complex has only modest effects on the nicotinamide ring of NADP+, but trimethoprim produces a substantial change in its environment. The dissociation rate constant of NADP+ from a number of complexes was also determined by saturation transfer.  相似文献   

10.
Dihydrofolate reductase (DHFR) catalyzes the NADPH-dependent reduction of dihydrofolate to tetrahydrofolate and is essential for the synthesis of thymidylate, purines and several amino acids. Inhibition of the enzyme's activity leads to arrest of DNA synthesis and cell death. The enzyme has been studied extensively as a drug target for bacterial, protozoal and fungal infections, and also for neoplastic and autoimmune diseases. Here, we report the crystal structure of dihydrofolate reductase from Mycobacterium tuberculosis, a human pathogen responsible for the death of millions of human beings per year. Three crystal structures of ternary complexes of M. tuberculosis DHFR with NADP and different inhibitors have been determined, as well as the binary complex with NADP, with resolutions ranging from 1.7 to 2.0 A. The three DHFR inhibitors are the anticancer drug methotrexate, the antimicrobial trimethoprim and Br-WR99210, an analogue of the antimalarial agent WR99210. Structural comparison of these complexes with human dihydrofolate reductase indicates that the overall protein folds are similar, despite only 26 % sequence identity, but that the environments of both NADP and of the inhibitors contain interesting differences between the enzymes from host and pathogen. Specifically, residues Ala101 and Leu102 near the N6 of NADP are distinctly more hydrophobic in the M. tuberculosis than in the human enzyme. Another striking difference occurs in a region near atoms N1 and N8 of methotrexate, which is also near atom N1 of trimethoprim, and near the N1 and two methyl groups of Br-WR99210. A glycerol molecule binds here in a pocket of the M. tuberculosis DHFR:MTX complex, while this pocket is essentially filled with hydrophobic side-chains in the human enzyme. These differences between the enzymes from pathogen and host provide opportunities for designing new selective inhibitors of M. tuberculosis DHFR.  相似文献   

11.
Dihydrofolate reductase from Lactobacillus casei and its complexes with NADPH and methotrexate yield well-resolved Raman spectra. The 1685-cm?1 Raman band assigned to the carboxamide of NADPH persists in the NADPH-enzyme binary complex but is absent from the NADPH-methotrexate-enzyme ternary complex. This is ascribed to stabilization of the polarized form of the carboxamide by H bonding to the NH and CO groups of Ala 6 and Ile 13 of the peptide backbone.  相似文献   

12.
R67 dihydrofolate reductase (DHFR) is a novel bacterial protein that possesses 222 symmetry and a single active site pore. Although the 222 symmetry implies that four symmetry-related binding sites must exist for each substrate as well as for each cofactor, various studies indicate only two molecules bind. Three possible combinations include two dihydrofolate molecules, two NADPH molecules, or one substrate plus one cofactor. The latter is the productive ternary complex. To explore the role of various ligand substituents during binding, numerous analogues, inhibitors, and fragments of NADPH and/or folate were used in both isothermal titration calorimetry (ITC) and K(i) studies. Not surprisingly, as the length of the molecule is shortened, affinity is lost, indicating that ligand connectivity is important in binding. The observed enthalpy change in ITC measurements arises from all components involved in the binding process, including proton uptake. As a buffer dependence for binding of folate was observed, this likely correlates with perturbation of the bound N3 pK(a), such that a neutral pteridine ring is preferred for pairwise interaction with the protein. Of interest, there is no enthalpic signal for binding of folate fragments such as dihydrobiopterin where the p-aminobenzoylglutamate tail has been removed, pointing to the tail as providing most of the enthalpic signal. For binding of NADPH and its analogues, the nicotinamide carboxamide is quite important. Differences between binary (binding of two identical ligands) and ternary complex formation are observed, indicating interligand pairing preferences. For example, while aminopterin and methotrexate both form binary complexes, albeit weakly, neither readily forms ternary complexes with the cofactor. These observations suggest a role for the O4 atom of folate in a pairing preference with NADPH, which ultimately facilitates catalysis.  相似文献   

13.
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   

14.
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   

15.
Structural and thermodynamic interactions for the binding of trimethoprim and related congeners to the binary complex of dihydrofolate reductase (from chicken) and NADPH are explored using free energy simulation methods. Good agreement between structures from experimental X-ray refinement and molecular dynamics simulations is found for the complexes. Agreement with thermodynamic measurements is found as well. Our thermodynamic calculations suggest that entropic contributions and desolvation thermodynamics can play a crucial role in overall binding, and that extreme care must be taken in the use of simple model building to rationalize or predict protein-drug binding.  相似文献   

16.
We have prepared a selectively deuterated dihydrofolate reductase in which all the aromatic protons except the C(2) protons of tryptophan have been replaced by deuterium and have examined the 1H NMR spectra of its complexes with folate, trimethoprim, methotrexate, NADP+, and NADPH. One of the four Trp C(2)-proton resonance signals (signal P at 3.66 ppm from dioxane) has been assigned to Trp-21 by examining the NMR spectrum of a selectively deuterated N-bromosuccinimide-modified dihydrofolate reductase. This signal is not perturbed by NADPH, indicating that the coenzyme is not binding close to the 2 position of Trp-21. This contrasts markedly with the 19F shift (2.7 ppm) observed for the 19F signal of Trp-21 in the NADPH complex with the 6-fluorotryptophan-labeled enzyme. In fact the crystal structure of the enzyme . methotrexate . NADPH shows that the carboxamide group of the reduced nicotinamide ring is near to the 6 position of Trp-21 but remote from its 2 position. The nonadditivity of the 1H chemical-shift contributions for signals tentatively assigned to Trp-5 and -133 indicates that these residues are influenced by ligand-induced conformational changes.  相似文献   

17.
The u.v. difference spectra generated when methotrexate, trimethoprim or folate bind to Lactobacillus casei dihydrofolate reductase were analysed. The difference spectrum producted by methotrexate binding is shown to consist of three components: (a) one closely resembling that observed on protonation of methotrexate, reflecting an increased degree of protonation on binding; (b) a pH-independent contribution corresponding to a 40 nm shift to longer wavelengths of a single absorption band of methotrexate: (c) a component arising from perturbation of tryptophan residue(s) of the enzyme. Quantitative analysis of the pH-dependence of component (a) shows that pK of methotrexate is increased from 5.35 to 8.55 (+/-0.10) on binding. In contrast, folate is not protonated when bound to the enzyme at neutral pH. At pH7.5, where methotrexate is bound 2000 times more tightly than folate, one-third of the difference in binding energy between the two compounds arises from the difference in chaarge stage. A similar analysis of the difference spectra generated on trimethoprim binding demonstrates that this compound, too, shows an increase in pK on binding but only from 7.22 to 7.90 (+/-0.10), suggesting that its 2,4-diaminopyrimidine ring does not bind to the enzyme in precisely the same way as the corresponding moiety of methotrexate.  相似文献   

18.
In order to examine the origins of the large positive cooperativity (ΔG(0)(coop) = -2.9 kcal mol(-1)) of trimethoprim (TMP) binding to a bacterial dihydrofolate reductase (DHFR) in the presence of NADPH, we have determined and compared NMR solution structures of L. casei apo DHFR and its binary and ternary complexes with TMP and NADPH and made complementary thermodynamic measurements. The DHFR structures are generally very similar except for the A-B loop region and part of helix B (residues 15-31) which could not be directly detected for L. casei apo DHFR because of line broadening from exchange between folded and unfolded forms. Thermodynamic and NMR measurements suggested that a significant contribution to the cooperativity comes from refolding of apo DHFR on binding the first ligand (up to -0.95 kcals mol(-1) if 80% of A-B loop requires refolding). Comparisons of Cα-Cα distance differences and domain rotation angles between apo DHFR and its complexes indicated that generally similar conformational changes involving domain movements accompany formation of the binary complexes with either TMP or NADPH and that the binary structures are approaching that of the ternary complex as would be expected for positive cooperativity. These favorable ligand-induced structural changes upon binding the first ligand will also contribute significantly to the cooperative binding. A further substantial contribution to cooperative binding results from the proximity of the bound ligands in the ternary complex: this reduces the solvent accessible area of the ligand and provides a favorable entropic hydrophobic contribution (up to -1.4 kcal mol(-1)).  相似文献   

19.
The binding of NADP+ to dihydrofolate reductase (EC 1.5.1.3) in the presence and absence of substrate analogs has been studied using 1H and 13C nuclear magnetic resonance (NMR). NADP+ binds strongly to the enzyme alone and in the presence of folate, aminopterin, and methotrexate with a stoichiometry of 1 mol of NADP+/mol of enzyme. In the 13C spectra of the binary and ternary complexes, separate signals were observed for the carboxamide carbon of free and bound [13CO]NADP+ (enriched 90% in 13C). The 13C signal of the NADP+-reductase complex is much broader than that in the ternary complex with methotrexate because of exchange line broadening on the binary complex signal. From the difference in line widths (17.5 +/- 3.0 Hz) an estimate of the dissociation rate constant of the binary complex has been obtained (55 +/- 10 sec-1). The dissociation rate of the NADP+-reductase complex is not the rate-limiting step in the overall reaction. In the various complexes studied large 13C chemical shifts were measured for bound [13CO]NADP+ relative to free NADP+ (upfield shifts of 1.6-4.3 ppm). The most likely origin of the bound shifts lies in the effects on the shieldings of electric fields from nearby charged groups. For the NADP+-reductase-folate system two 13C signals from bound NADP+ are observed indicating the presence of more than one form of the ternary complex. The IH spectra of the binary and ternary complexes confirm both the stoichiometry and the value of the dissociation rate constant obtained from the 13C experiments. Substantial changes in the IH spectrum of the protein were observed in the different complexes and these are distinct from those seen in the presence of NADPH.  相似文献   

20.
S R Stone  J F Morrison 《Biochemistry》1988,27(15):5493-5499
Kinetic studies on the reaction catalyzed by dihydrofolate reductase from Escherichia coli have been undertaken with the aim of characterizing further the kinetic mechanism of the reaction. For this purpose, the kinetic properties of substrates were determined by measurement of (a) initial velocities over a wide range of substrate concentrations and (b) the stickiness of substrates in ternary enzyme complexes. Stickiness is defined as the rate at which a substrate reacts to give products relative to the rate at which that substrate dissociates. Stickiness was determined by varying the viscosity of reaction mixtures and the concentration of one substrate in the presence of a saturating concentration of the other substrate. The results indicate that NADPH is sticky in the enzyme-NADPH-dihydrofolate complex, while dihydrofolate is much less sticky in this complex. At higher concentrations, NADPH functions as an activator through the formation of an enzyme-NADPH-tetrahydrofolate from which tetrahydrofolate is released more rapidly than from an enzyme-tetrahydrofolate complex. Higher concentrations of dihydrofolate also cause enzyme activation, and it appears that this effect is due to the ability of dihydrofolate to displace tetrahydrofolate from a binary enzyme complex through the formation of a transitory enzyme-tetrahydrofolate-dihydrofolate complex. As NADPH and dihydrofolate function as activators and as NADPH behaves as a sticky substrate, the kinetic mechanism of the dihydrofolate reductase reaction with the natural substrates is steady-state random. By contrast with NADPH, reduced 3-acetylpyridine adenine dinucleotide phosphate exhibits only slight stickiness and does not function as an activator.(ABSTRACT TRUNCATED AT 250 WORDS)  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