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 78 毫秒
1.
The effects of alpha-, beta- and gamma-cyclodextrins on the amylose and maltopentaose hydrolysis catalysed by porcine pancreatic alpha-amylase (PPA) were investigated. The results of the statistical analysis performed on the kinetic data using the general initial velocity equation of a one-substrate reaction in the presence of one inhibitor indicate that the type of inhibition involved depends on the substrate used: the inhibition of amylose hydrolysis by alpha-, beta- and gamma-cyclodextrin is of the competitive type, while the inhibition of maltopentaose hydrolysis is of the mixed noncompetitive type. Consistently, the Lineweaver-Burk plots intersect on the vertical axis when amylose is used as the substrate, while in the case of maltopentaose, the intersection occurs at a point located in the second quadrant. The inhibition of the hydrolysis therefore involves only one abortive complex, PPA-cyclodextrin, when amylose is used as the substrate, while two abortive complexes, PPA-cyclodextrin and PPA-maltopentaose-cyclodextrin, are involved with maltopentaose. The mixed noncompetitive inhibition thus shows the existence of one accessory binding site. In any case, only one molecule of inhibitor binds to PPA. In line with these findings, the difference spectra of PPA produced by alpha-, beta- and gamma-cyclodextrin indicate that binding occurs at a tryptophan and a tyrosine residue. The corresponding dissociation constants and the inhibition constants obtained using the kinetic approach are in the same range (1.2-7 mM). The results obtained here on the inhibition of maltopentaose hydrolysis by cyclodextrin are similar to those previously obtained with acarbose as the inhibitor [Alkazaz, M., Desseaux, V., Marchis-Mouren, G., Prodanov, E. & Santimone, M. (1998) Eur. J. Biochem. 252, 100-107], but differ from those obtained with amylose as the substrate and acarbose as inhibitor [Alkazaz, M., Desseaux, V., Marchis-Mouren, G., Payan, F., Forest, E. & Santimone, M. (1996) Eur. J. Biochem. 241, 787-796]. It is concluded that the hydrolysis of both long and short chain substrates requires at least one secondary binding site, including a tryptophan residue.  相似文献   

2.
Two inhibitors, acarbose and cyclodextrins (CD), were used to investigate the active site structure and function of barley alpha-amylase isozymes, AMY1 and AMY2. The hydrolysis of DP 4900-amylose, reduced (r) DP18-maltodextrin and maltoheptaose (catalysed by AMY1 and AMY2) was followed in the absence and in the presence of inhibitor. Without inhibitor, the highest activity was obtained with amylose, kcat/Km decreased 103-fold using rDP18-maltodextrin and 10(5) to 10(6)-fold using maltoheptaose as substrate. Acarbose is an uncompetitive inhibitor with inhibition constant (L1i) for amylose and maltodextrin in the micromolar range. Acarbose did not bind to the active site of the enzyme, but to a secondary site to give an abortive ESI complex. Only AMY2 has a second secondary binding site corresponding to an ESI2 complex. In contrast, acarbose is a mixed noncompetitive inhibitor of maltoheptaose hydrolysis. Consequently, in the presence of this oligosaccharide substrate, acarbose bound both to the active site and to a secondary binding site. alpha-CD inhibited the AMY1 and AMY2 catalysed hydrolysis of amylose, but was a very weak inhibitor compared to acarbose.beta- and gamma-CD are not inhibitors. These results are different from those obtained previously with PPA. However in AMY1, as already shown for amylases of animal and bacterial origin, in addition to the active site, one secondary carbohydrate binding site (s1) was necessary for activity whereas two secondary sites (s1 and s2) were required for the AMY2 activity. The first secondary site in both AMY1 and AMY2 was only functional when substrate was bound in the active site. This appears to be a general feature of the alpha-amylase family.  相似文献   

3.
The general unireactant modifier mechanism in the absence of product can be described by the following linked reactions: E + S k1 in equilibrium k-1 ES k3----E + P; E + I k5 in equilibrium k-5 EI; EI + S k2 in equilibrium k-2 ESI k4----EI + P; and ES + I k6 in equilibrium k-6 ESI where S is a substrate and I is an effector. A full steady state treatment yields a velocity equation that is second degree in both [S] and [I]. Two different conditions (or assumptions) permit reduction of the velocity equation to one that is first degree in [S] and [I]. These are (a) that k-2k3 = k-1k4 (Frieden, C., J. Biol. Chem. 239, pp. 3522-3531, (1964)) and (b) that the I-binding reactions are at equilibrium (Reinhart, G. D., Arch. Biochem. Biophys. 224, pp. 389-401 (1983)). It is shown that each condition gives rise to the other (i.e., if the I-binding reactions are at equilibrium, then k-2k3 must equal k-1k4 and vice-versa). If one assumes equilibrium for the I-binding steps, the velocity equation derived by the method of Cha (J. Biol. Chem. 243, pp. 820-825 (1968)) is apparently second degree in [I] (Segel, I. H., Enzyme Kinetics, p. 838, Wiley-Interscience (1975)), but reduces to a first degree equation when the relationship derived by Frieden is inserted. If one starts by assuming a single equilibrium condition for I binding, e.g., k-5[EI] = k5[E][I] or k-6[ESI] = k6[ES][I], then a traditional algebraic manipulation of the remaining steady state equations provides first degree expressions for the concentrations of all enzyme species and also discloses the Frieden relationship.  相似文献   

4.
White kidney bean (Phaseolus vulgaris) contains two kinds of alpha-amylase inhibitors, one heat-stable (alpha AI-s) and one heat-labile (alpha AI-u). alpha AI-s has recently been revealed to be a tetrameric complex, alpha(2)beta(2), with two active sites [Kasahara et al. (1996) J. Biochem. 120, 177-183]. The present study was undertaken to reveal the molecular features of alpha AI-u, which is composed of three kinds of subunits, alpha, beta, and gamma. The gamma-subunit, in contrast to the alpha- and beta-subunits that are indistinguishable from the alpha- and beta-subunits of alpha AI-s, was found to correspond to a subunit of an alpha-amylase inhibitor-like protein, which has been identified as an inactive, evolutionary intermediate between arcelin and the alpha-amylase inhibitor in a P. vulgaris defense protein family. The polypeptide molecular weight of alpha AI-u determined by the light-scattering technique, together with the polypeptide molecular weights of the subunits, suggests that alpha AI-u is a trimeric complex, alpha beta gamma. The inhibition of alpha AI-u by increasing amounts of porcine pancreatic alpha-amylase (PPA) indicates that an inactive 1:1 complex is formed between alpha AI-u and PPA. Molecular weight estimation of the complex by the light-scattering technique confirmed that it is a complex of alpha AI-u with one PPA molecule. Thus it seems probable that alpha AI-u is an evolutionary intermediate of the P. vulgaris alpha-amylase inhibitor.  相似文献   

5.
The inhibition of highly purified rat liver L-threonine dehydratase (L-threonine hydro-lyase (deaminating), EC 4.2.1.16) by aminothiols (L-cysteine, D-cysteine, cysteamine) has been studied. Single inhibition experiments evaluated by Lineweaver-Burk and Dixon plots showed, in a given concentration range, partially (parabolic) competitive inhibitions, indicating two binding sites for each inhibitor. Double inhibition experiments revealed that the inhibition was antagonistic, the two inhibitors weakening each other's effect. Formation of EI1 and EI2 binary complexes, and ESI1, ESI2 and EI1I2 ternary complexes was demonstrated, while formation of the quaternary complex ESI1I2 was ruled out. It is assumed that one inhibitor-binding site coincides with the substrate-binding center while the second inhibitor-binding (allosteric, regulatory) site may comprise the pyridoxal-phosphate-binding SH group(s). The comparison between Km and Ki values and the evaluation of intracellular concentrations of L-threonine, L-cysteine and cysteamine suggest a possible physiological role of the inhibition.  相似文献   

6.
Limited proteolysis of the pyruvate decarboxylase (E1, alpha2beta2) component of the pyruvate dehydrogenase (PDH) multienzyme complex of Bacillus stearothermophilus has indicated the importance for catalysis of a site (Tyr281-Arg282) in the E1alpha subunit (Chauhan, H.J., Domingo, G.J., Jung, H.-I. & Perham, R.N. (2000) Eur. J. Biochem. 267, 7158-7169). This site appears to be conserved in the alpha-subunit of heterotetrameric E1s and multiple sequence alignments suggest that there are additional conserved amino-acid residues in this region, part of a common pattern with the consensus sequence -YR-H-D-YR-DE-. This region lies about 50 amino acids on the C-terminal side of a 30-residue motif previously recognized as involved in binding thiamin diphosphate (ThDP) in all ThDP-dependent enzymes. The role of individual residues in this set of conserved amino acids in the E1alpha chain was investigated by means of site-directed mutagenesis. We propose that particular residues are involved in: (a) binding the 2-oxo acid substrate, (b) decarboxylation of the 2-oxo acid and reductive acetylation of the tethered lipoyl domain in the PDH complex, (c) an "open-close" mechanism of the active site, and (d) phosphorylation by the E1-specific kinase (in eukaryotic PDH and branched chain 2-oxo acid dehydrogenase complexes).  相似文献   

7.
This study is the first report on the effectiveness and specificity of alpha-acarviosinyl-(1-->4)-alpha-D-glucopyranosyl-(1-->6)-D-glucopyranosylidene-spiro-thiohydantoin (PTS-G-TH) inhibitor on the 2-chloro-4-nitrophenyl-4-O-beta-D-galactopyranosyl-maltoside (GalG2CNP) and amylose hydrolysis catalysed by human salivary alpha-amylase (HSA). Synthesis of PTS-G-TH was carried out by transglycosylation using acarbose as donor and glucopyranosylidene-spiro-thiohydantoin (G-TH) as acceptor. This new compound was found to be a much more efficient HSA inhibitor than G-TH. The inhibition is a mixed-noncompetitive type on both substrates and only one molecule of inhibitor binds to the enzyme. Kinetic constants calculated from secondary plots are in micromolar range. Values of K(EI) and K(ESI) are very similar in the presence of GalG2CNP substrate; 0.19 and 0.24 microM, respectively. Significant difference can be found for K(EI) and K(ESI) using amylose as substrate; 8.45 and 0.5 microM, respectively. These values indicate that inhibition is rather uncompetitive than competitive related to amylose hydrolysis.  相似文献   

8.
Pyruvate has previously been shown to slow down the rate of intramolecular electron transfer from the flavosemiquinone (Fs) to the cytochrome b2 moiety of flavocytochrome b2 [Tegoni, M., Silvestrini, M. C., Labeyrie, F. & Brunori, M. (1984) Eur. J. Biochem. 140, 39-45] and to stabilize markedly the Fs state of the prosthetic flavin, relative to the oxidized (Fo) and the reduced (Fh) states [Tegoni, M., Janot, J. M. & Labeyrie, F. (1986) Eur. J. Biochem. 155, 491-503]. In the present study, we have determined the dissociation constants of pyruvate for the three redox forms of the prosthetic flavin and demonstrated that the Fs-pyruvate complex is actually much more stable than the Fo-pyruvate and Fh-pyruvate complexes. The inhibition produced by pyruvate has been characterized under steady-state conditions using both ferricytochrome c and ferricyanide as external acceptor. A detailed analysis and simulations of the suitable reaction scheme, taking into consideration all data from rapid kinetic studies of partial reactions previously published, show that the experimental noncompetitive inhibition results from the sum of a competitive effect due to binding of pyruvate to Fo and an uncompetitive effect due to binding to the Fs intermediate in a dead-end complex. Pyruvate binding to the semiquinone transient results in a marked loss of the reactivity of this donor in electron transfers to its specific partner, the cytochrome b2 present in the same active site, as to ferricyanide, an external acceptor. A critical evaluation of the parameters involved in the control of such reactivities is presented.  相似文献   

9.
The rapidly reacting cysteine-sulfhydryl group of L-isoleucyl-tRNA synthetase has been specifically alkylated with L-isoleucyl-bromomethyl ketone [Rainey, P., Holler, E. & Kula, M.-R. (1976) Eur. J. Biochem. 63, 419-426]. We have now investigated the catalytic and substrate binding properties of the modified protein by radioactive and fluorescence techniques. The rate constants for the transfer of AMP and isoleucine from the protein - adenylate complex to form ATP or Ile-tRNAIle were only 3% of those for native enzyme, whereas the rate constant for the formation of adenylate was essentially unchanged. The tendency to form synthetase - substrate complexes remained almost unchanged with the exception of L-isoleucine which exhibited a 20-fold reduction. Similarly, complex formation of L-isoleucinol together with its synergistic coupling to complex formation of ATP was partially inhibited. The results rule out the essential participation of the rapidly alkylatable cysteine-sulfhydryl group during catalysis.  相似文献   

10.
C A Gates  D B Northrop 《Biochemistry》1988,27(10):3826-3833
Aminoglycoside nucleotidyltransferase 2'-I conveys multiple antibiotic resistance to Gram-negative bacteria because the enzyme adenylylates a broad range of aminoglycoside antibiotics as substrates [Gates, C. A., & Northrop, D. B. (1988) Biochemistry (preceding paper in this issue)]. The enzyme also catalyzes the transfer of a variety of nucleotides [Van Pelt, J. E., & Northrop, D. B. (1984) Arch. Biochem. Biophys. 230, 250-263]. This doubly broad substrate specificity makes it an excellent candidate for application of the alternative substrate diagnostic [Radika, K., & Northrop, D. B. (1984) Anal. Biochem. 141, 413-417] as a means to determine its kinetic mechanism. The kinetic patterns presented here are composed of one set of intersecting lines and one coincident line and are consistent with a Theorell-Chance kinetic mechanism in which nucleotide binding precedes aminoglycosides, pyrophosphate is released prior to the nucleotidylated aminoglycoside (Q), and turnover is controlled by the rate-limiting release of the final product. Substrate inhibition by tobramycin (B) is partial and uncompetitive versus Mg-ATP, indicating that B binds to the EQ complex, but not in the usual dead-end fashion common to an ordered sequential release of products; instead, Q may escape from the abortive EQB complex at a finite rate. Dead-end inhibition by neomycin C (I) is also partial and uncompetitive versus Mg-ATP but is slope-linear, intercept-hyperbolic, partial noncompetitive versus gentamicin A; both kinetic patterns signify the formation of a partial abortive EQI complex.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The effectiveness and specificity of a tannin inhibition on human salivary amylase (HSA) catalyzed hydrolysis was studied using 2-chloro-4-nitrophenyl 4-O-beta-D-galactopyranosyl-alpha-maltoside (GalG(2)-CNP) and amylose substrates. Aleppo tannin was isolated from the gall nut of Aleppo oak. This tannin is a gallotannin, in which glucose is esterified with gallic acids. This is the first kinetic report, which details the inhibitory effects of this compound on HSA. A mixed non-competitive type inhibition has been observed on both substrates. The extent of inhibition is markedly dependent on the substrate-type. Kinetic constants were calculated from Lineweaver-Burk secondary plots for GalG(2)-CNP (K(EI) 0.82 microg mL(-1), K(ESI) 3.3 microg mL(-1)). This indicates a 1:1 binding ratio of inhibitor-enzyme and/or inhibitor-enzyme-substrate complex. When amylose was the substrate the binding ratio of inhibitor to enzyme-substrate complex was found to be 2:1, with the binding constants of K(EI) 17.4 microg mL(-1), K(ESI) 14.9 microg mL(-1), K(ESI(2)) 9.6 microg mL(-1). Presumably, the tannin inhibitor can bind not only to HSA, but to the amylose substrate, as well. Kinetic data suggest that Aleppo tannin is a more efficient amylase inhibitor than the recently studied other tannin with quinic acid core (GalG(2)-CNP: K(EI) 9.0 microg mL(-1), K(ESI) 47.9 microg mL(-1)).  相似文献   

12.
The extent of insertion of beta-strand s4A into sheet A in intact serpin alpha 1-proteinase inhibitor (alpha 1PI has been probed by peptide annealing experiments [Schulze et al. (1990) Eur. J. Biochem. 194, 51-56]. Twelve synthetic peptides of systematically varied length corresponding in sequence to the unprimed (N-terminal) side of the active site loop were complexed with alpha 1PI. The complexes were then characterized by circular dichroism spectroscopy and tested for inhibitory activity. Four peptides formed complexes which retained inhibitory activity, one of which was nearly as effective as the native protein. Comparison with the three dimensional structures of cleaved alpha 1PI [L?bermann et al. (1984) J. Mol. Biol. 177, 531-556] and plakalbumin [Wright et al. (1990) J. Mol. Biol. 213, 513-528] supports a model in which alpha 1PI requires the insertion of a single residue, Thr345, into sheet A for activity.  相似文献   

13.
Triclinic crystals of the complex formed by eglin with subtilisin Carlsberg were analyzed by X-ray diffraction. The crystal and molecular structure of this complex was determined with data that extended to 0.12-nm resolution by a combination of Patterson search methods and isomorphous replacement techniques. Its structure was refined to a crystallographic R value of 0.178 (1.0-0.12 nm) using an energy-restraint least-squares procedure. The complete subtilisin molecule could be traced without ambiguity in the refined electron density. The eglin component, from which an amino-terminal segment is cleaved off, is only defined from Lys8I (i.e. the lysine residue 8 of the inhibitor) onwards. Per unit cell, 436 fixed solvent molecules and 2 calcium ions were located. In spite of 84 amino acid replacements and one deletion, subtilisin Carlsberg exhibits a very similar polypeptide fold to subtilisin BPN'. The root-mean-square deviations of all alpha-carbon atoms (excluding those at the deletion site) from models of subtilisin BPN' [Alden, R. A., Birktoft, J. J., Kraut, J., Robertus, J. D. & Wright, C. S. (1971) Biochem. Biophys. Res. Commun. 45, 337-344] and subtilisin Novo [Drenth, J., Hol, W. G. J., Jansonius, J. N. & Kockoek, R. (1972) Eur. J. Biochem. 25, 177-181] are 0.077 nm and 0.103 nm. Most of these deviations result from global shifts rather than changes of the local geometry. The single-residue deletion at position 56 affects only the surrounding conformation. Two sites of high electron density and close distances to surrounding oxygen ligands have been found in the Carlsberg enzyme which are probably occupied by calcium ions. Eglin consists of a twisted four-stranded beta-sheet flanked by an alpha-helix and by an exposed proteinase binding loop on opposite sides. Around the reactive site, Leu45I-Asp46I, this loop is mainly stabilized by electrostatic/hydrogen bond interactions with the side chains of two arginine residues which project from the hydrophobic core [Bode, W., Papamokos, E., Musil, D., Seemüller, W. & Fritz, H. (1986) EMBO J. 5, 813-818]. The reactive site loop conformation resembles that found in other 'small' proteinase inhibitors. The scissile peptide bond is not cleaved but its carbonyl group is slightly distorted from planar geometry. Most of the intermolecular contacts are contributed by the nine residues of the reactive-site loop Gly40I-Arg48I.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
In mammals, carbamoyl phosphate for utilization in pyrimidine biosynthesis is synthesized by a glutamine-dependent carbamoyl-phosphate synthase II which is subject to regulation by 5-phospho-alpha-D-ribosyl 1-diphosphate (PRib-PP), a positive effector, and MgUTP, a negative effector [Mori, M., Ishida, H. and Tatibana, M. (1975) Biochemistry 14, 2622-2630]. We have found that Lineweaver-Burk plots of carbamoyl phosphate synthase activity versus 1/[MgATP] are described by a velocity equation which is a ratio of quadratic polynomials, consistent with a positive homotropic interaction between two catalytic sites for the binding of MgATP (Ks = 16.6 +/- 3.1 mM, interaction factor a = 0.00538 +/- 0.00245). The activating effect of PRib-PP upon carbamoyl-phosphate synthase is consistent with PRib-PP binding at an allosteric site (Ka = 31.4 +/- 6.4 microM) and promoting the binding of a first molecule of MgATP as substrate (interaction factor l = 0.0437 +/- 0.0063). Thus MgATP and PRib-PP bind to the E X MgATP complex with respective dissociation constants of a X Ks = 0.089 mM and l X Ka = 1.4 microM while MgATP binds to the E X PRib-PP complex with a dissociation constant of l X Ks = 0.73 mM. Data for the inhibitory effect of MgUTP upon carbamoyl-phosphate synthase indicate that MgUTP competes with MgATP for binding at the catalytic site (Ki = 0.203 +/- 0.016 mM). A computer model has recently been developed which enables quantitative stimulation of the time-dependent effects of blockade of the pyrimidine pathway by a tight-binding enzyme inhibitor [Duggleby, R.G. and Christopherson, R.I. (1984) Eur. J. Biochem. 143, 221-226]. The velocity equation derived in the present paper provides a quantitative basis for predicting changes in the flux through the de novo pyrimidine pathway in growing cells.  相似文献   

15.
Elongation factor EF-Tu (Mr approximately equal to 50 000) and elongation factor EF-G (Mr approximately equal to 78 000) were isolated from Bacillus stearothermophilus in a homogeneous form. The ability of EF-Tu to participate in protein synthesis is rapidly inactivated by N-tosyl-L-phenyl-alanylchloromethane (Tos-PheCH2Cl). EF-Tu X GTP is more susceptible to the inhibition by Tos-PheCH2Cl than is EF-Tu X GDP. Tos-PheCH2Cl forms a covalent equimolar complex with the factor by reacting with a cysteine residue in its molecule. The labelling of EF-Tu by the reagent irreversibly destroys its ability to bind aminoacyl-tRNA, which in turn protects the protein from this inactivation. This indicates that the modification of EF-Tu by Tos-PheCH2Cl occurs at the aminoacyl-tRNA binding site of the protein. To identify and characterize the site of aminoacyl-tRNA binding in EF-Tu, the factor was labelled with [14C]Tos-PheCH2Cl, digested with trypsin, the resulting peptides were separated by high-performance liquid chromatography and the sequence of the radioactive peptide was determined. The peptide has identical structure with an Escherichia coli EF-Tu tryptic peptide comprising the residues 75-89 and the Tos-PheCH2Cl-reactive cysteine at position 81 [Jonák, J., Petersen, T. E., Clark, B. F. C. and Rychlík, I. (1982) FEBS Lett. 150, 485-488]. Experiments on photo-oxidation of EF-Tu by visible light in the presence of rose bengal dye showed that there are apparently two histidine residues in elongation factor Tu from B. stearothermophilus which are essential for the interaction with aminoacyl-tRNA. This is clearly reminiscent of a similar situation in E. coli EF-Tu [Jonák, J., Petersen, T. E., Meloun, B. and Rychlík, I. (1984) Eur. J. Biochem. 144, 295-303]. Our results provide further evidence for the conserved nature of the site of aminoacyl-tRNA binding in elongation factor EF-Tu and show that Tos-PheCH2Cl reagent might be a favourable tool for the identification of the site in the structure of prokaryotic EF-Tus.  相似文献   

16.
The paramagnetic effects of the bound manganese ion and of a covalently attached spin label on proton nuclear spin relaxation rates have been used to calculate distances for a structural model of the MnADP and creatine complexed to creatine kinase from rabbit muscle. The nucleotide and guanidino substrates are so aligned on the enzyme that the transferable phosphoryl group on one substrate is in apposition to the acceptor moiety on the second substrate. The divalent metal ion is most probably liganded to the alpha and beta phosphates of the nucleotide substrate, both in the abortive MnADP-creatine-enzyme complex and in the active MnATP-creatine-enzyme complex. The metal ion-formate distance approximately 5 A in the Mn(II)ADP-formate-creatine-enzyme complex and less than 5 A in the Co(II)ADP-formate-creatine-enzyme complex is consistent with the suggestion that the monovalent anion is binding at the site normally occupied by the transferable phosphoryl group, thus producing a complex which mimics the transition state. Although only an upper limit of the distance from Mn(II) to the guanidino substrate could be determined in the presence of formate, it could be concluded that the disposition of the guanidino substrate changes upon addition of formate, since the relative distances of the methyl and methylene group are inverted. The effect of formate and nitrate on increasing the residence time of creatine in the MnADP-creatine-enzyme complex as determined by NMR provides evidence that the complexes observed by NMR are identical with those involved in the catalytic mechanism, since a parallel effect of formate and nitrate is observed in the kinetics of the enzymatic reaction, where the dissociation constant of creatine from the abortive quaternary complex decreases in the presence of the anions as had been determined from their inhibition of the forward reaction (Milner-White, E.J., and Watts, D.C. (1971) Biochem. J. 122, 727-740). Although the guanidino substrate is not directly liganded to the divalent metal ion, the electron paramagnetic resonance spectrum of manganese in the transition state analog complexes, i.e. nitrate-ADP-guanidino substrate-enzyme, is strongly dependent on catalytic activity of the guanidino substrate. The structural differences observed by EPR among transition state analog complexes with various guanidino substrates were not reflected in distances from Mn(II) to the guanidino substrate, which were 10% and 0.3% as active as creatine. Within the experimental error of 1 A, the distances were the same. The enzyme or the enzyme-substrate complexes may be considered to exist in a number of structurally distinct conformations in equilibrium based on the EPR spectra and on the anomalous temperature-dependence of the relaxation rates of the formate proton of the transition state analog complexes...  相似文献   

17.
Fries M  Jung HI  Perham RN 《Biochemistry》2003,42(23):6996-7002
Pyruvate decarboxylase (E1) catalyzes the first two reactions of the four involved in oxidative decarboxylation of pyruvate by the pyruvate dehydrogenase (PDH) multienzyme complex. It requires thiamin diphosphate to bring about the decarboxylation of pyruvate, which is followed by the reductive acetylation of a lipoyl group covalently bound to the N(6) amino group of a lysine residue in the second catalytic component, a dihydrolipoyl acetyltransferase (E2). Replacement of two histidine residues in the E1alpha and E1beta chains of the heterotetrameric E1 (alpha(2)beta(2)) component of the PDH complex of Bacillus stearothermophilus, considered possible proton donors at the active site, was carried out. Subsequent characterization of the mutants permitted different roles to be assigned to these two particular residues in the reaction catalyzed by E1: E1alpha His271 to stabilize the dianion formed during decarboxylation of the 2-oxo acid and E1beta His128 to provide the proton required to protonate the incoming dithiolane ring in the subsequent reductive acetylation of the lipoyl goup. On the basis of these and other results from a separate investigation into the roles of individual residues in a loop region in the E1alpha chain close to the active site of E1 [Fries, M., Chauhan, H. J., Domingo, G. J., Jung, H., and Perham, R. N. (2002) Eur. J. Biochem. 270, 861-870] together with work from other laboratories, a detailed mechanism for the E1 reaction can be formulated.  相似文献   

18.
Phosphopyridoxyl derivatives, which are stable analogues of a substrate-coenzyme complex, are bound at the active site with great affinity. From a comparison of the interaction of a number of such compounds with the apoenzyme the delta G0 values for the binding of the substrate carboxy and phenyl groups and of the coenzyme aldehydic group were determined to be equal to (or more negative than) -3.8. -8.4 and -12.5kJ/mol (-0.9, -1.9 and -3kcal/mol) respectively; the delta G0 for the binding of the coenzyme phosphate group was shown to be more negative than -20.5kJ/mol (-4.9kcal/mol). Two features of the binding process of the coenzyme-substrate analogues to tyrosine decarboxylase have already been found in the case of tyrosine aminotransferase [Borri-Voltattorni, Orlacchio, Giartosio, Conti & Turano (1975) Eur. J. Biochem. 53, 151-160]: (1) in the binding of the substrate to the enzyme a significant fraction of the instrinsic delta G0 appears to be used for some associated endoergonic process; (2) the delta H0 and delta S0 of binding appear to be very sensitive indicators of the correct alignment of the substrate-coenzyme and analogues at the active site.  相似文献   

19.
Kinetically, the three-step model proposed for the interaction between beta-lactam antibiotics and the exocellular DD-carboxypeptidases-transpeptidases of Streptomyces R61 and Actinomadura R39 [Frère, Ghuysen & Iwatsubo (1975) Eur. J. Biochem. 57, 343--357; Fuad, Frère, Ghuysen, Duez & Iwatsubo (1976) Biochem. J. 155, 623--629] applies to the interaction between the much less penicillin-sensitive exocellular DD-carboxypeptidase-endopeptidase of Streptomyces albus G and at least phenoxymethylpenicillin, cephalothin and cephalosporin C. The penicillin resistance of the albus G enzyme is mainly due to the low efficiency with which the first reversible complex formed with the antibiotic (complex EI) undergoes transformation into a second more stable complex EI*. Analysis of the ternary interaction between enzyme, NalphaNepsilon-diacetyl-L-lysyl-D-alanyl-D-alanine (Ac2-L-Lys-D-Ala-D-Ala) and cephalosporin C indicates a non-competitive mechanism.  相似文献   

20.
A thermodynamic model is presented to describe the redox behaviour of the tetraheme cytochrome c3 from Desulfovibrio gigas. This molecule displays different intrinsic redox potentials for the four hemes and during the redox titration process, interactions among different hemes occur, thus altering the values of redox potentials according to which of the hemes are oxidized [Santos, H., Moura, J.J.G., Moura, I., LeGall, J. & Xavier, A.V. (1984) Eur. J. Biochem. 141, 283-296]. This complex cooperative behaviour [Xavier, A.V. (1986) J. Inorg. Biochem. 28, 239-243] has been analyzed here using an I2H4-interaction network [Cornish-Bowden, A. & Koshland, D.E. Jr (1970) J. Biol. Chem. 245, 6241-6250] coupled to a proton-linked equilibrium between two tertiary structures. Such a formalism, which requires a reduced number of parameters, is able to fully account quantitatively for the pH dependence of the NMR redox-titration curves. The 'redox-Bohr' effect is discussed in terms of the available structure and thermodynamic data and a functional mechanism is proposed.  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