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 31 毫秒
1.
The three-dimensional structures of cystatins, and other evidence, suggest that the flexible N-terminal region of these inhibitors may bind to target proteinases independent of the two rigid hairpin loops forming the remainder of the inhibitory surface. In an attempt to demonstrate such two-step binding, which could not be identified in previous kinetics studies, we introduced a cysteine residue before the N-terminus of cystatin A and labeled this residue with fluorescent probes. Binding of AANS- and AEDANS-labeled cystatin A to papain resulted in approximately 4-fold and 1.2-fold increases of probe fluorescence, respectively, reflecting the interaction of the N-terminal region with the enzyme. Observed pseudo-first-order rate constants, measured by the loss of papain activity in the presence of a fluorogenic substrate, for the reaction of the enzyme with excess AANS-cystatin A increased linearly with the concentration of the latter. In contrast, pseudo-first-order rate constants, obtained from measurements of the change of probe fluorescence with either excess enzyme or labeled inhibitor, showed an identical hyperbolic dependence on the concentration of the reactant in excess. This dependence demonstrates that the binding occurs in two steps, and implies that the labeled N-terminal region of cystatin A interacts with the proteinase in the second step, subsequent to the hairpin loops. The comparable affinities and dissociation rate constants for the binding of labeled and unlabeled cystatin A to papain indicate that the label did not appreciably perturb the interaction, and that unlabeled cystatin therefore also binds in a similar two-step manner. Such independent binding of the N-terminal regions of cystatins to target proteinases after the hairpin loops may be characteristic of most cystatin-proteinase reactions.  相似文献   

2.
The reduction of gluconolactone by glucose-fructose oxidoreductase containing tightly bound NADPH (enzyme-NADPH) is biphasic in nucleotide fluorescence. The initial rapid decrease, which represents quenching of the fluorescence by bound lactone, is followed by a slower decrease which corresponds to the change in absorbance. At low glucose concentrations, the oxidation of glucose by enzyme-NADP+ involves a single first-order process with similar rate constants in fluorescence and absorbance. At higher glucose concentrations, the apparent first-order rate constants for the fluorescence change are less than those for the absorbance change. This is consistent with a mechanism in which the fluorescence change occurs during the lactone dissociation step, which is slower than the hydrogen transfer step during which the absorbance change occurs. The rate constant for gluconolactone dissociation is 360 +/- 10 s-1 and this step is therefore rate-determining for the overall reaction. Reduction of fructose by enzyme-NADPH is first order with a limiting rate constant of at least 2000 s-1.  相似文献   

3.
The rate constants for both acylation and deacylation of beta-lactamase PC1 from Staphylococcus aureus and the RTEM beta-lactamase from Escherichia coli were determined by the acid-quench method [Martin & Waley (1988) Biochem. J. 254, 923-925] with several good substrates, and, for a wider range of substrates, of beta-lactamase I from Bacillus cereus. The values of the acylation and deacylation rate constants for benzylpenicillin were approximately the same (i.e. differing by no more than 2-fold) for each enzyme. The variation of kcat./Km for benzylpenicillin with the viscosity of the medium was used to obtain values for all four rate constants in the acyl-enzyme mechanism for all three enzymes. The reaction is partly diffusion-controlled, and the rate constant for the dissociation of the enzyme-substrate complex has approximately the same value as the rate constants for acylation and deacylation. Thus all three first-order rate constants have comparable values. Here there is no single rate-determining step for beta-lactamase action. This is taken to be a sign of a fully efficient enzyme.  相似文献   

4.
Porter DJ  Short SA 《Biochemistry》2000,39(38):11788-11800
The catalytically active form of monofunctional yeast orotidine-5'-phosphate decarboxylase was a dimer (E(2)). The dimer equilibrium dissociation constant was 0.25 microM in 0.01 M MOPS Na(+) at pH 7.2. The bimolecular rate constant for dimer formation was 1.56 microM(-1) s(-1). The dimeric form of the enzyme was stabilized by NaCl such that the enzyme was E(2) in 100 mM NaCl at all concentrations of enzyme tested. The kinetics of binding of OMP to E(2) was governed by two ionizations (pK(1) = 6.1 and pK(2) = 7.7). From studies with substrate analogues, the higher pK was assigned to a group on the enzyme that interacted with the pyrimidinyl moiety. The value of the lower pK was dependent on the substrate analogue, which suggested that it was not exclusively the result of ionization of the phosphoryl moiety. During the decarboxylation of OMP, the fluorescence of E(2) was quenched over 20%. The enzymatic species with reduced fluorescence was a catalytically competent intermediate that had kinetic properties consistent with it being the initial enzyme-substrate complex. The stoichiometry for binding of OMP to E(2) was one OMP per enzyme monomer. The value of the first-order rate constant for conversion of the enzyme-substrate complex to free enzyme (36 s(-1)) calculated from a single turnover experiment ([E] > [S]) was slightly greater than the value of k(cat), 20 s(-1) (corrected for stoichiometry), calculated from steady-state data. In the single turnover experiments, the enzyme was E(2)*S, whereas in the steady-state turnover the experiment enzyme was E(2)*S(2). The similarity of these values suggested that the subunits were catalytically independent such that E(2)*S(2) could be treated as E*S and that conversion of the enzyme-substrate complex to E was k(cat). Kinetic data for the approach to the steady-state with OMP and E(2) yield a bimolecular association rate complex of 62 microM(-1) s(-1)and a dissociation rate constant for E*S of 60 s(-1). The commitment to catalysis was 0.25. By monitoring the effect of carbonic anhydrase on [H(+)] changes during a single turnover experiment, the initial product of the decarboxylation reaction was shown to be CO(2) not HCO(3-). UMP was released from the enzyme concomitantly with CO(2) during the conversion of E*S to E. Furthermore, the enzyme removed an enzyme equivalent of H(+) from solvent during this step of the reaction. The bimolecular rate constants for association of 6-AzaUMP and 8-AzaXMP, substrate analogues with markedly different nucleobases, had association rate constants of 112 and 130 microM(-1) s(-1), respectively. These results suggested that the nucleobase did not contribute significantly to the success of formation of the initial enzyme-substrate complex.  相似文献   

5.
Dihydrofolate reductase from a methotrexate-resistant subline (R6) of L1210 mouse leukemia cells is activated (i.e. has its catalytic activity increased severalfold) by treatment with (a) sulfhydryl-modifying agents (p-chloromercuribenzoate (pCMB) or 5,5'-dithiobis(2-nitrobenzoic acid], (b) salts (KCl or NaCl), or (c) chaotropes (urea or guanidinium hydrochloride). With b or c activation is rapid (less than 10 s), but with a the process is much slower; at 25 degrees C, pseudo first-order rate constants for activation by excess pCMB or 5,5'-dithiobis(2-nitrobenzoic acid) are 0.45 and 0.08 min-1, respectively. Activation can also be monitored by conformational changes in the protein as indicated by enhanced fluorescence of 2-p-toluidinylnaphthalene-6-sulfonate or by increased intrinsic fluorescence of tryptophan residues in the enzyme. Pseudo first-order rate constants for the pCMB-induced conformational change, measured by these fluorimetric procedures (0.45 min-1 and about 0.4 min-1, respectively), are in good agreement with the value obtained from the increase in catalytic activity. The rate of modification of the single cysteine residue in the enzyme by excess 14C-labeled pCMB, however, is faster than the rate of activation, indicating that the conformational change follows derivatization and is the rate-limiting step in the overall process. Activated forms of the enzyme are more labile to thermal denaturation or proteolysis than the untreated enzyme; the former process, however, is retarded by the presence of bovine serum albumin. Activation by the various agents is considered to involve a common mechanism in which interaction of the enzyme with the agents is followed by conformational changes in the enzyme, producing a series of forms that differ in microstructure, catalytic activity, and lability.  相似文献   

6.
Solubilization of excess sludge (ES) cultivated in a laboratory and that obtained from real wastewater treatment plants by continuous Dyno Mill disruption were studied for the purpose of effective reuse of ES. The total soluble protein concentration released during the ES disruption was found to be closely dependent upon the released protease activity, which suggests the possibility of the enzyme recovery as one of the ES reuse methods. The operating parameters of Dyno Mill significantly affected the ES solubilization. A high ES solubilization ratio of about 53%, defined by soluble TOC per total carbon contained in the laboratory-cultivated ES, was achieved by using glass beads of 0.5mm in diameter with the 60% (v/v) bead loading at the disc agitating speed of 3200rpm. The ES solubilization by the continuous mill disruption was successfully analyzed by a kinetic model consisting of first-order disruption kinetics. The collision frequency between the moving beads correlated well with the ES solubilization rate constant under all the operating conditions. This study proposes the applicability of mill disruption to the ES solubilization and to its reuse process.  相似文献   

7.
The oxidation of a series of primary alcohols by liver alcohol dehydrogenase has been studied under conditions of [S]o greater than [E]o using the stopped-flow method. A biphasic process, with exponential rise to a steady state, was observed for most of the alcohols and the rate constants for the transient phase were determined. No transient phase could be detected for 2-chloroethanol and 2-nitroethanol and steady-state measurements were made for these alcohols. The rate constants for the hydrogen transfer step were obtained from the pre-steady-state rate constants for the various alcohols and correlated with the Taft sigma constant. The (see article) value obtained (-1.8) is consistent with rate-limiting hydride transfer coupled with removal of the hydroxyl proton by a suitable basic group on the enzyme. A possible identity for this group is suggested.  相似文献   

8.
Purified adenosine kinase from L1210 cells displayed substrate inhibition by high concentrations of adenosine (Ado), ATP, and MgCl2. When incubated with ATP and MgCl2, the enzyme was phosphorylated, and the phosphorylated kinase transferred phosphate to adenosine in the absence of ATP and MgCl2. Substrate binding, isotope exchange, and kinetic studies suggested that the enzyme catalyzes the reaction by means of a two-site ping-pong mechanism with the phosphorylated enzyme as an obligatory intermediate. Among many possible pathways within this mechanism probably a random-bi ordered-bi route is the preferred sequence in which the two substrates, adenosine and MgATP, bind in a random order to form the ternary complex MgATP . E . Ado followed by the sequential dissociation of MgADP and AMP. Dissociation constants of various enzyme-substrate and enzyme-product complexes and the first-order rate constant of the rate-limiting step were estimated.  相似文献   

9.
The 3'-processing of viral DNA extremities is the first step in the integration process catalysed by human immunodeficiency virus (HIV)-1 integrase (IN). This reaction is relatively inefficient and processed DNAs are usually detected in vitro under conditions of excess enzyme. Despite such experimental conditions, steady-state Michaelis-Menten formalism is often applied to calculate characteristic equilibrium/kinetic constants of IN. We found that the amount of processed product was not significantly affected under conditions of excess DNA substrate, indicating that IN has a limited turnover for DNA cleavage. Therefore, IN works principally in a single-turnover mode and is intrinsically very slow (single-turnover rate constant = 0.004 min(-1)), suggesting that IN activity is mainly limited at the chemistry step or at a stage that precedes chemistry. Moreover, fluorescence experiments showed that IN-DNA product complexes were very stable over the time-course of the reaction. Binding isotherms of IN to DNA substrate and product also indicate tight binding of IN to the reaction product. Therefore, the slow cleavage rate and limited product release prevent or greatly reduce subsequent turnover. Nevertheless, the time-course of product formation approximates to a straight line for 90 min (apparent initial velocity), but we show that this linear phase is due to the slow single-turnover rate constant and does not indicate steady-state multiple turnover. Finally, our data ruled out the possibility that there were large amounts of inactive proteins or dead-end complexes in the assay. Most of complexes initially formed were active although dramatically slow.  相似文献   

10.
To better understand substrate recognition and catalysis by RNase III, we examined steady-state and pre-steady-state reaction kinetics, and changes in intrinsic enzyme fluorescence. The multiple turnover cleavage of a model RNA substrate shows a pre-steady-state burst of product formation followed by a slower phase, indicating that the steady-state reaction rate is not limited by substrate cleavage. RNase III catalyzed hydrolysis is slower at low pH, permitting the use of pre-steady-state kinetics to measure the dissociation constant for formation of the enzyme-substrate complex (K(d)=5.4(+/-0.6) nM), and the rate constant for phosphodiester bond cleavage (k(c)=1.160(+/-0.001) min(-1), pH 5.4). Isotope incorporation analysis shows that a single solvent oxygen atom is incorporated into the 5' phosphate of the RNA product, which demonstrates that the cleavage step is irreversible. Analysis of the pH dependence of the single turnover rate constant, k(c), fits best to a model for two or more titratable groups with pK(a) of ca 5.6, suggesting a role for conserved acidic residues in catalysis. Additionally, we find that k(c) is dependent on the pK(a) value of the hydrated divalent metal ion included in the reaction, providing evidence for participation of a metal ion hydroxide in catalysis, potentially in developing the nucleophile for the hydrolysis reaction. In order to assess whether conformational changes also contribute to the enzyme mechanism, we monitored intrinsic tryptophan fluorescence. During a single round of binding and cleavage by the enzyme we detect a biphasic change in fluorescence. The rate of the initial increase in fluorescence was dependent on substrate concentration yielding a second-order rate constant of 1.0(+/-0.1)x10(8) M(-1) s(-1), while the rate constant of the second phase was concentration independent (6.4(+/-0.8) s(-1); pH 7.3). These data, together with the unique dependence of each phase on divalent metal ion identity and pH, support the hypothesis that the two fluorescence transitions, which we attribute to conformational changes, correlate with substrate binding and catalysis.  相似文献   

11.
The formation and dissociation of the aldolase-dihydroxyacetone phosphate complex were studied by following changes in A240 [Topper, Mehler & Bloom (1957), Science 126, 1287-1289]. It was shown that the enzyme-substrate complex (ES) slowly isomerizes according to the following reaction: (formula: see text) the two first-order rate constants for the isomerization step being k+2 = 1.3s-1 and k-2 = 0.7s-1 at 20 degrees C and pH 7.5. The dissociation of the ES complex was provoked by the addition of the competitive inhibitor hexitol 1,6-bisphosphate. At 20 degrees C and pH 7.5, k+1 was 4.7 X 10(6)M-1-S-1 and k-1 was 30s-1. Both the ES and the ES* complexes react rapidly with 1.7 mM-glyceraldehyde 3-phosphate, the reaction being practically complete in 40 ms. This shows that the ES* complex is not a dead-end complex. Evidence was also provided that aldolase binds and utilizes only the keto form of dihydroxyacetone phosphate.  相似文献   

12.
A series of 1,4-naphthoquinone derivatives diversely substituted at C-2, C-3, C-5 and C-8, prepared by reaction of amines, amino acids and alcohols with commercial 1,4-naphthoquinones, has been evaluated against papain and bovine spleen cathepsin B. These 1,4-naphthoquinone derivatives were found to be irreversible inhibitors for both cysteine proteases, with second-order rate constants, k(2), ranging from 0.67 to 35.4M(-1)s(-1) for papain, and from 0.54 to 8.03M(-1)s(-1) for cathepsin B. Some derivatives display a hyperbolic dependence of the first-order inactivation rate constant, k(obs), with the inhibitor concentration, indicative of a specific interaction process between enzyme and inhibitor. The chemical reactivity of the compounds towards cysteine as a model thiol is dependent on the naphthoquinone LUMO energy, whereas papain inactivation is not. The 1,4-naphthoquinone derivatives are inactive against the serine protease, porcine pancreatic elastase.  相似文献   

13.
Lignin peroxidase compound III. Mechanism of formation and decomposition   总被引:9,自引:0,他引:9  
Lignin peroxidase compound III (LiPIII) was prepared via three procedures: (a) ferrous LiP + O2 (LiPIIIa), (b) ferric LiP + O2-. (LiPIIIb), and (c) LiP compound II + excess H2O2 followed by treatment with catalase (LiPIIIc). LiPIIIa, b, and c each have a Soret maximum at approximately 414 nm and visible bands at 543 and 578 nm. LiPIIIa, b, and c each slowly reverted to native ferric LiP, releasing stoichiometric amounts of O2-. in the process. Electronic absorption spectra of LiPIII reversion to the native enzyme displayed isosbestic points in the visible region at 470, 525, and 597 nm, suggesting a single-step reversion with no intermediates. The LiPIII reversion reactions obeyed first-order kinetics with rate constants of approximately 1.0 X 10(-3) s-1. In the presence of excess peroxide, at pH 3.0, native LiP, LiPII, and LiPIIIa, b, and c are all converted to a unique oxidized species (LiPIII*) with a spectrum displaying visible bands at 543 and 578 nm, but with a Soret maximum at 419 nm, red-shifted 5 nm from that of LiPIII. LiPIII* is bleached and inactivated in the presence of excess H2O2 via a biphasic process. The fast first phase of this bleaching reaction obeys second-order kinetics, with a rate constant of 1.7 X 10(1) M-1 s-1. Addition of veratryl alcohol to LiPIII* results in its rapid reversion to the native enzyme, via an apparent one-step reaction that obeys second-order kinetics with a rate constant of 3.5 X 10(1) M-1 s-1. Stoichiometric amounts of O2-. are released during this reaction. When this reaction was run under conditions that prevented further reactions, HPLC analysis of the products demonstrated that veratryl alcohol was not oxidized. These results suggest that the binding of veratryl alcohol to LiPIII* displaces O2-., thus returning the enzyme to its native state. In contrast, the addition of veratryl alcohol to LiPIII did not affect the rate of spontaneous reversion of LiPIII to the native enzyme.  相似文献   

14.
Pressure, as a perturbing variable, is one of the most powerful tools to investigate the thermodynamic parameters of chemical reactions and to study the mechanism of enzyme-catalyzed reactions. The effect of elevated hydrostatic pressure (up to 0.8 kbar) on the reaction of butyrylcholinesterase with N-methyl-(7-dimethylcarbamoxy)quinolinium was determined under single-turnover conditions at 35 degrees C. The rate of carbamylation was monitored as the accumulation of the fluorescent ion, N-methyl-7-hydroxyquinolinium, in a high-pressure stopped-flow apparatus designed for the assay of fluorescence. Elevated pressure favored formation of the enzyme-substrate complex but inhibited carbamylation of the enzyme. Because a single reaction step was recorded, it was possible to interpret the data obtained under high pressure in the form of Michaelis-Menten equations. From the pressure dependence of the dissociation constant for the enzyme-substrate complex and the rate constant for carbamylation, maximal volume changes accompanying these events were determined. The value for the binding process, delta Vb = -129 ml.mol-1, is too large to be related only to volumetric changes in the active center. Substrate-induced conformational change and change of water structure appear to be the dominant contributions to the overall volume change associated with substrate binding. The large positive activation volume measured (delta V not equal to = 119 ml.mol-1) may also reflect extended structural and hydration changes. At pressures greater than 0.4 kbar, an additional pressure effect, dependent on substrate concentration, occurred in a narrow pressure interval. This effect may have resulted from a substrate-induced pressure-sensitive enzyme conformational state.  相似文献   

15.
E W Taylor 《Biochemistry》1977,16(4):732-739
The transient phase of adenosine triphosphate (ATP) hydrolysis (early burst) was investigated for myosin, heavy meromyosin (HMM), and subfragment 1 (S-1) over a range of temperatures and pH's. The burst size at pH 8,20 degrees C, is 0.8-0.85, based on steady-state and transient measurements. The equilibrium constant for the enzyme-substrate to enzyme-product transition is 0.85 +/- 0.05. It is concluded that both myosin heads undergo the rapid hydrolysis step and that there are no significant differences for S-1 vs. HMM or myosin. The transient data are fitted reasonably well by a single rate process, but available evidence is consistent with some heterogeneity and a range of rate constants differing by a factor of two. At pH 6.9 and 3 degrees C, the burst size is 0.5 and the hydrolysis is slower than the configuration change measured by fluorescence. The results are consistent with the kinetic scheme (see article). The lower burst at low temperature and pH can be partly explained by a reduction in the equilibrium constant, K3, and ATP can be synthesized on the enzyme by a pH-temperature jump.  相似文献   

16.
1. The rate equation for a generalized Michaelian type of enzymic reaction mechanism has been analyzed in order to establish how the mechanism should be kinetically designed in order to optimize the catalytic efficiency of the enzyme for a given average magnitude of true and apparent first-order rate constants in the mechanism at given concentrations of enzyme, substrate and product. 2. As long as on-velocity constants for substrate and product binding to the enzyme have not reached the limiting value for a diffusion-controlled association process, the optimal state of enzyme operation will be characterized by forward (true and apparent) first-order rate constants of equal magnitude and reverse rate constants of equal magnitude. The drop in free energy driving the catalysed reaction will occur to an equal extent for each reaction step in the mechanism. All internal equilibrium constants will be of equal magnitude and reflect only the closeness of the catalysed reaction to equilibrium conditions. 3. When magnitudes of on-velocity constants for substrate and product binding have reached their upper limits, the optimal kinetic design of the reaction mechanism becomes more complex and has to be established by numerical methods. Numerical solutions, calculated for triosephosphate isomerase, indicate that this particular enzyme may or may not be considered to exhibit close to maximal efficiency, depending on what value is assigned to the upper limit for a ligand association rate constant. 4. Arguments are presented to show that no useful information on the evolutionary optimization of the catalytic efficiency of enzymes can be obtained by previously taken approaches that are based on the application of linear free-energy relationships for rate and equilibrium constants in the reaction mechanism.  相似文献   

17.
Dimeric ligands can be potent inhibitors of protein-protein or enzyme-substrate interactions. They have increased affinity and specificity toward their targets due to their ability to bind two binding sites simultaneously and are therefore attractive in drug design. However, few studies have addressed the kinetic mechanism of interaction of such bivalent ligands. We have investigated the binding interaction of a recently identified potent plasma-stable dimeric pentapeptide and PDZ1–2 of postsynaptic density protein-95 (PSD-95) using protein engineering in combination with fluorescence polarization, isothermal titration calorimetry, and stopped-flow fluorimetry. We demonstrate that binding occurs via a two-step process, where an initial binding to either one of the two PDZ domains is followed by an intramolecular step, which produces the bidentate complex. We have determined all rate constants involved in the binding reaction and found evidence for a conformational transition of the complex. Our data demonstrate the importance of a slow dissociation for a successful dimeric ligand but also highlight the possibility of optimizing the intramolecular association rate. The results may therefore aid the design of dimeric inhibitors in general.  相似文献   

18.
It is a common practice to employ k cat[E]0/K m as a first-order rate constant for the analysis of an enzymatic reaction, where [E]0 is the total enzyme concentration. I describe in this report a serious shortcoming in analyzing enzymatic reactions when k cat[E]0/K m is employed and show that k cat[E]0/K m can only be applied under very limited conditions. I consequently propose the use of a more universal first-order rate constant, k cat[ES]K/[S]0, where [ES]K is the initial equilibrium concentration of the ES-complex derived from [E]0, [S]0 and K m. Employing k cat[ES]K/[S]0 as the first-order rate constant enables all enzymatic reactions to be reasonably simulated under a wide range of conditions, and the catalytic and binding contributions to the rate constant of any enzyme can be determined under any and all conditions.  相似文献   

19.
Stopped-flow techniques were utilized to investigate the kinetics of the reaction of lignin peroxidase compounds I and II (LiPI and LiPII) with veratryl alcohol (VA). All rate data were collected from single turnover experiments under pseudo first-order conditions. The reaction of LiPI with VA strictly obeys second-order kinetics over the pH range 2.72-5.25 as demonstrated by linear plots of the pseudo first-order rate constants versus concentrations of VA. The second-order rate constants are strongly dependent on pH and range from 2.62 x 10(6) M-1 s-1 (pH 2.72) to 1.45 x 10(4) M-1 s-1 (pH 5.25). The reaction of LiPII and VA exhibits saturation behavior when the observed pseudo first-order rate constants are plotted against VA concentrations. The saturation phenomenon is quantitatively explained by the formation of a 1:1 LiPII-substrate complex. Results of kinetic and rapid scan spectral analyses exclude the formation of a LiPII-VA cation radical complex. The first-order dissociation rate constant and the equilibrium dissociation constant for the LiPII reaction are also pH dependent. Binding of VA to LiPII is controlled by a heme-linked ionizable group of pKa approximately 4.2. The pH profiles of the second-order rate constants for the LiPI reaction and of the first-order dissociation constants for the LiPII reaction both demonstrate two pKa values at approximately 3.0 and approximately 4.2. Protonated oxidized enzyme intermediates are most active, suggesting that only electron transfer, not proton uptake from the reducing substrate, occurs at the enzyme active site. These results are consistent with the one-electron oxidation of VA to an aryl cation radical by LiPI and LiPII.  相似文献   

20.
Kinetic specificity in papain-catalysed hydrolyses   总被引:1,自引:12,他引:1       下载免费PDF全文
The specificity of the proteolytic enzyme, papain, for the peptide bond of the substrate adjacent to that about to be cleaved and for the acyl residue of some N-acylglycine derivatives is manifest almost exclusively in the formation of the acyl-enzyme from the enzyme-substrate complex. Models for the enzyme-substrate complex and acyl-enzyme intermediate are suggested that account for these observations. In particular it is suggested that the peptide bond of the substrate adjacent to that about to be cleaved, is bound in the cleft of the enzyme between the NH group of glycine-66 and the backbone C=O group of aspartic acid-158, and provides a sensitive amplification mechanism through which the specificity of the enzyme for hydrophobic amino acids such as l-phenylalanine is relayed. It is also suggested that the distortion in the enzyme-substrate complex and the binding of the peptide bond adjacent to that about to be cleaved are also linked and behave co-operatively, the distortion of the protein facilitating binding and the stronger binding facilitating distortion. The results imply that between the enzyme-substrate complex and the acyl-enzyme a relaxation of the protein conformation must occur.  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