首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The ionization state of the phosphate group bound at the aspartate aminotransferase apoenzyme's active site has been investigated utilizing Fourier-transform infrared spectroscopy following the band corresponding to the symmetric stretching of the dianionic phosphate. Unlike free phosphate, when inorganic phosphate is bound at the enzyme's active site, the integrated intensity value of the dianionic band does not change with pH within the studied range, and this value is similar to that for free dianionic phosphate at pH 8.3. From these results, we propose a dianionic state for the phosphate ion bound to cytosolic aspartate aminotransferase throughout the pH range of 5.7-8.3. The presence of other anions such as acetate and chloride or the substrate aspartate and its analogues produces a pH-dependent phosphate removal from the active site which is favored at low pH values. Elimination of the charged primary amine at the active-site Lys-258, through formation of a Schiff base with pyridoxal or chemical modification by carbamylation, also produces a pH-independent phosphate release. These results are interpreted as Lys-258 together with the active-site alpha-helix and other residues may be involved in stabilizing phosphate as a dianion in the apoenzyme phosphate pocket which anchors the phosphate ester of pyridoxal phosphate in the holoenzyme. It is proposed that the dianionic phosphate contributes to the apoenzyme's thermal stability through formation of strong hydrogen bond and salt bridges with the amino acid residues forming the phosphate binding pocket with assistance of Lys-258, and other active-site cationic components.  相似文献   

2.
The crystal structures of four inhibitor complexes of dialkylglycine decarboxylase are reported. The enzyme does not undergo a domain closure, as does aspartate aminotransferase, upon inhibitor binding. Two active-site conformations have been observed in previous structures that differ in alkali metal ion content, and two active-site conformations have been shown to coexist in solution when a single type of metal ion is present. There is no indication of coexisting conformers in the structures reported here or in the previously reported structures, and the observed conformation is that expected based on the presence of potassium in the enzyme. Thus, although two active-site conformations coexist in solution, a single conformation, corresponding to the more active enzyme, predominates in the crystal. The structure of 1-aminocyclopropane-1-carboxylate bound in the active site shows the aldimine double bond to the pyridoxal phosphate cofactor to be fully out of the plane of the coenzyme ring, whereas the Calpha-CO2(-) bond lies close to it. This provides an explanation for the observed lack of decarboxylation reactivity with this amino acid. The carboxylate groups of both 1-aminocyclopropane-1-carboxylate and 5'-phosphopyridoxyl-2-methylalanine interact with Ser215 and Arg406 as previously proposed. This demonstrates structurally that alternative binding modes, which constitute substrate inhibition, occur in the decarboxylation half-reaction. The structures of d and l-cycloserine bound to the active-site show that the l-isomer is deprotonated at C(alpha), presumably by Lys272, while the d-isomer is not. This difference explains the approximately 3000-fold greater potency of the l versus the d-isomer as a competitive inhibitor of dialkylglycine decarboxylase.  相似文献   

3.
The phenolic group of active site residue Tyr-248 in carboxypeptidase A has a pKa value of 10.06, as determined from the pH dependence of its rate of nitration by tetranitromethane. The decrease in enzyme activity (kcat/Km) in alkaline solution, characterized by a pKa value of approximately 9.0 (for cobalt carboxypeptidase A), is associated with the protonation state of an imidazole ligand of the active-site metal ion, as indicated by a selective pH dependence of the 1H NMR spectrum of the enzyme. Inhibition of the cobalt-substituted enzyme by 2-(1-carboxy-2-phenylethyl)phenol and its 4,6-dichloro- and 4-phenylazo-derivatives confirms that the decrease in enzyme activity (kcat/Km) in acidic solution, characterized by a pKa value of 5.8, is due to the protonation state of a water molecule bound to the active-site metal ion in the absence of substrate. Changes in the coordination number of the active-site metal ion are seen in its visible absorption spectrum as a consequence of binding of the phenolic inhibitors. Conventional concepts regarding the mechanisms of the enzyme are brought into question.  相似文献   

4.
Deoxyhypusine synthase catalyzes the first step in the two-step post-translational synthesis of hypusine, which is uniquely present in eukaryotic initiation factor 5A (eIF5A). Deoxyhypusine synthase and eIF5A are conserved throughout the eukaryotic kingdom, and both are essential for cell proliferation and survival. A previous study (Liao, D. I., Wolff, E. C., Park, M. H., and Davies, D. R. (1998) Structure 6, 23-32) of human deoxyhypusine synthase revealed four active sites of the homotetrameric enzyme located within deep tunnels. These Form I crystals were obtained under conditions of acidic pH and high ionic strength and likely contain an inactive enzyme. Each active-site entrance is blocked by a ball-and-chain motif composed of a region of extended structure capped by a two-turn alpha-helix. We report here at 2.2 A a new Form II crystal of the deoxyhypusine synthase:NAD holoenzyme grown at low ionic strength and pH 8.0, near the optimal pH for enzymatic activity. The ball-and-chain motif could not be detected in the electron density, suggesting that it swings freely and thus it no longer obstructs the active-site entrance. The deoxyhypusine synthase competitive inhibitor N(1)-guanyl-1,7-diaminoheptane (GC(7))is observed bound within the putative active site of the enzyme in the new crystal form (Form II) after exposure to the inhibitor. This first structure of a deoxyhypusine synthase.NAD.inhibitor ternary complex under physiological conditions now provides a structural context to discuss the results of previous biochemical investigations of the deoxyhypusine synthase reaction mechanism. This structure also provides a basis for the development of improved inhibitors and antiproliferative agents.  相似文献   

5.
H J Dyson  L L Tennant  A Holmgren 《Biochemistry》1991,30(17):4262-4268
A series of two-dimensional (2D) correlated 1H NMR spectra of reduced and oxidized Escherichia coli thioredoxin have been used to probe the effects of pH in the vicinity of the active site, -Cys32-Gly-Pro-Cys35-, using the complete proton resonance assignments available for thioredoxin. In either oxidation state, the majority of residues of the thioredoxin molecule remain unchanged between pH 5.7 and pH 10, as indicated by the identical chemical shifts of the C alpha H, C beta H, and other protons. In reduced thioredoxin, a fairly widespread region around the active-site dithiol is affected by the titration of a group or groups with pKa approximately 7.1-7.4 in 2H2O. Another titration, with pKa approximately 8.4, affects a smaller region of the protein. Oxidized thioredoxin contains a disulfide and no free thiol groups; nevertheless, the proton resonances of many groups in the active-site region were observed to titrate with a pKa of 7.5, probably as a result of an abnormally high pKa value for the carboxyl group of the buried Asp-26 residue. For reduced thioredoxin, the results indicate that Asp-26 is titrating in this pH range, as well as both thiol groups. The new results are strongly suggestive that the mechanism of thioredoxin-catalyzed protein disulfide reduction may be critically dependent on proton transfer as well as electron transfer within the active site.  相似文献   

6.
Rabbit muscle pyruvate kinase was irreverisbly inactivated by 5-chloro-4-oxopentanoic acid with a pKa of 9.2. The inhibition was time-dependent and was related to the 5-chloro-4-oxopentanoic acid concentration. Analysis of the kinetics of inhibition showed that the binding of the inhibitor showed positive co-operativity (n = 1.5 +/- 0.2). Inhibition of pyruvate kinase by 5-chloro-4-oxopentanoic acid was prevented by ligands which bind to the active site. Their effectiveness was placed in the order Mg2+ greater than phosphoenolpyruvate greater than ATP greater than ADP greater than pyruvate. Inhibitor-modified pyruvate kinase was unable to catalyse the detritiation of [3-(3)H]pyruvate in the ATP-promoted reaction, but it did retain 5-10% of the activity with either phosphate or arsenate as promoters. 5-Chlor-4-oxo-[3,5-(3)H]pentanoic acid was covalently bound to pyruvate kinase and demonstrated a stoicheiometry of 1 mol of inhibitor bound per mol of pyruvate kinase protomer. The incorporation of the inhibitor and the loss of enzyme was proportional. These results are discussed in terms of 5-chloro-4-oxopentanoic acid alkylating a functional group in the phosphoryl overlap region of the active site, and a model is presented in which this compound alkylates an active-site thiol in a reaction that is controlled by a more basic group at the active site.  相似文献   

7.
An extensive structural manifold of short hydrogen bond-mediated, active site-directed, serine protease inhibition motifs is revealed in a set of over 300 crystal structures involving a large suite of small molecule inhibitors (2-(2-phenol)-indoles and 2-(2-phenol)-benzimidazoles) determined over a wide range of pH (3.5-11.4). The active site hydrogen-bonding mode was found to vary markedly with pH, with the steric and electronic properties of the inhibitor, and with the type of protease (trypsin, thrombin or urokinase type plasminogen activator (uPA)). The pH dependence of the active site hydrogen-bonding motif is often intricate, constituting a distinct fingerprint of each complex. Isosteric replacements or minor substitutions within the inhibitor that modulate the pK(a) of the phenol hydroxyl involved in short hydrogen bonding, or that affect steric interactions distal to the active site, can significantly shift the pH-dependent structural profile characteristic of the parent scaffold, or produce active site-binding motifs unique to the bound analog.Ionization equilibria at the active site associated with inhibitor binding are probed in a series of the protease-inhibitor complexes through analysis of the pH dependence of the structure and environment of the active site-binding groups involved in short hydrogen bond arrays. Structures determined at high pH (>11), suggest that the pK(a) of His57 is dramatically elevated, to a value as high as approximately 11 in certain complexes. K(i) values involving uPA and trypsin determined as a function of pH for a set of inhibitors show pronounced parabolic pH dependence, the pH for optimal inhibition governed by the pK(a) of the inhibitor phenol involved in short hydrogen bonds. Comparison of structures of trypsin, thrombin and uPA, each bound by the same inhibitor, highlights important structural variations in the S1 and active sites accessible for engineering notable selectivity into remarkably small molecules with low nanomolar K(i) values.  相似文献   

8.
Rabbit polyclonal antibody was raised to a chemically synthesized nonapeptide (Trp-Ala-Glu-Trp-Cys-Gly-Pro-Cys-Lys) corresponding to the active-site sequence of Escherichia coli thioredoxin. The antiserum efficiently inhibited thioredoxin activity in the standard thioredoxin reductase/NADPH coupled assay. This inhibition was blocked by preincubation of the antiserum with the nonapeptide. Tight association of the E. coli thioredoxin to the active-site antibody required SDS denaturation. These results suggest that thioredoxin reductase (NADPH: oxidized-thioredoxin oxidoreductase, EC 1.6.4.5) alters the conformation of thioredoxin sufficiently to permit binding to the antibody. The antiserum bound to plant and liver thioredoxins. Bovine pancreatic trypsin inhibitor, whose active site (Gly-Pro-Cys-Lys) is homologous to that of thioredoxin, also competes for the active-site antibody. This result led to experiments showing that thioredoxin can inhibit the digestion of cytochrome c by trypsin. The ability of thioredoxin to act as a trypsin inhibitor analogue provides a rationale for thioredoxin's resistance to digestion by trypsin.  相似文献   

9.
Endonuclease I of bacteriophage T7 is a DNA junction-resolving enzyme. We have previously used crystallography to demonstrate the binding of two manganese ions into the active site that is formed by three carboxylate (Glu 20, Asp 55 and Glu 65) and a lysine residue (Lys 67). Endonuclease I is active in the presence of magnesium, manganese, iron (II) and cobalt (II) ions, weakly active in the presence of nickel, copper (II) and zinc ions, and completely inactive in the presence of calcium ions. However, using calorimetry, we have observed the binding of two calcium ions to the free enzyme in a manner very similar to the binding of manganese ions. In the presence of iron (II) ions, we have obtained a cleavage of the continuous strands of a junction bound by endonuclease I, at sites close to (but not identical with) enzyme-induced hydrolysis. The results suggest that this arises from attack by locally generated hydroxyl radicals, arising from iron (II) ions bound into the active site. This therefore provides an indirect way of examining metal ion binding in the enzyme-junction complex. Ion binding in free protein (by calorimetry) and the enzyme-junction complex (iron-induced cleavage) have been studied in series of active-site mutants. Both confirm the importance of the three carboxylate ligands, and the lack of a requirement for Lys67 for the ion binding. Calorimetry points to particularly critical role of Asp55, as mutation completely abolishes all binding of both manganese and calcium ions.  相似文献   

10.
1. DL-alpha-Bromo-beta(5-imidazolyl)-propionic acid is a potential affinity labelling reagent for metallo-enzymes. It has been used with the alcohol dehydrogenases from liver and yeast. The liver enzyme is chemically modified and inactivated in a Michaelis-Menten-type reaction, where one molecule of the reagent is bound per subunit. The enzyme is protected from the inhibitor in a competitive manner by imidazole, 2,2'-dipyridyl, 1,10-phenanthroline and cyclohexanone, which all combine with the active-site zinc. The protection by chloride, acetate and NADH, which are considered to bind at the general anion binding site, is not strictly competitive. Inactivation has an optimum at pH 8.5. For the liver enzyme, the reagent was found to decrease the initial rate of ethanol oxidation. Prior to the irreversible alkylation of Cys-46, reversible binding is shown to occur at the active-site zinc atom. The yeast enzyme was extremely resistant to the reagent and no specific modification was found. 2. The potential affinity labelling and crosslinking reagent, symmetrical 1,3-dibromoacetone although unstable, has also been used for chemical modification. With the liver enzyme, concentrations below 5 mM gave a reaction of the Michaelis-Menten-type at pH 7.0. Several ligands known to complex with the active-site region protect the enzyme against the reagent. Dibromoacetone gave rapid inactivation of the yeast enzyme. Despite the fact that a pseudo-first-order reaction was observed with respect to enzyme as well as inhibitor, no saturating effect was found. In this work, dibromoacetone reacted like a monofunctional reagent.  相似文献   

11.
P-glycolate phosphatase requires divalent cations for activity. Activity-pH curves identified 2 active site residues with pK values at pH 5.7 and pH 9.1 in the presence of magnesium and at pH 5.7 and pH 7.5 in the presence of manganese or cobalt. Saturation velocity kinetics enabled the identification of two distinct divalent cation binding sites. The first, nonspecific site has a K0.5 of 2 to 7 x 10(-5) M, depending on the cation and the pH. The second site, which is specific for magnesium, binds this cation in a negatively cooperative fashion. The affinity at pH 8.1 varies approximately 100-fold from the first magnesium bound to the fourth. The negative cooperativity is greatest at high pH. Because the pH range of activity is very broad, both the phosphate monoanion and dianion of P-glycolate must be bound as the substrate. The concentration of these two species at the apparent Km is independent of magnesium concentration. The P-glycolate.magnesium complex is kinetically inactive.  相似文献   

12.
Metals bound to proteins perform a number of crucial biological reactions, including the oxidation of water by a manganese cluster in photosystem II. Although evolutionarily related to photosystem II, bacterial reaction centers lack both a strong oxidant and a manganese cluster for mediating the multielectron and proton transfer needed for water oxidation. In this study, carboxylate residues were introduced by mutagenesis into highly oxidizing reaction centers at a site homologous to the manganese-binding site of photosystem II. In the presence of manganese, light-minus-dark difference optical spectra of reaction centers from the mutants showed a lack of the oxidized bacteriochlorophyll dimer, while the reduced primary quinone was still present, demonstrating that manganese was serving as a secondary electron donor. On the basis of these steady-state optical measurements, the mutant with the highest-affinity site had a dissociation constant of approximately 1 microM. For the highest-affinity mutant, a first-order rate with a lifetime of 12 ms was observed for the reduction of the oxidized bacteriochlorophyll dimer by the bound manganese upon exposure to light. The dependence of the amplitude of this component on manganese concentration yielded a dissociation constant of approximately 1 muM, similar to that observed in the steady-state measurements. The three-dimensional structure determined by X-ray diffraction of the mutant with the high-affinity site showed that the binding site contains a single bound manganese ion, three carboxylate groups (including two groups introduced by mutagenesis), a histidine residue, and a bound water molecule. These reaction centers illustrate the successful design of a redox active metal center in a protein complex.  相似文献   

13.
The superoxide dismutases (EC 1.15.1.1) are a family of enzymes that catalyze the dismutation of superoxide radical anion to dioxygen and hydrogen peroxide. The active site contains a critical metal ion such as manganese, iron, or copper. The copper-containing protein also has one zinc ion bound per subunit. The standard method used to remove the metal ions from Cu,Zn superoxide dismutase has been to exhaustively dialyze the protein against chelating agents at low pH. We have developed a new method where the protein is bound to ion-exchange medium based on iminodiacetic acid immobilized on Sepharose. The bound protein is treated with a buffer containing edta at pH 3.5 to remove metal ions; the buffer is then exchanged for acetate buffer to remove edta, after which the protein is eluted by a salt gradient. An advantage of this method is that a single chromatography step is sufficient to produce apo protein. Results are shown for both human and bovine dimeric Cu,Zn superoxide dismutase and the monomeric Escherichia coli Cu,Zn superoxide dismutase. In every case, the metals were removed efficiently.  相似文献   

14.
The pH dependence of papain catalysis was analyzed by a scheme which evaluates the kinetic contribution of both protonated and unprotonated species of functional groups involved in catalysis. Kinetic measurements were made at constant pH, without buffers, by automatic titration. The rate-determining step for papain-catalyzed hydrolysis of alpha-N-benzoyl-L-arginine ethyl ester, determined by nucleophile competition, changed from acylation below pH 6.5 to mixed acylation-deacylation above pH 6.5. Kinetic analysis indicated that three prototropic groups governed the pH-specificity of alpha-N-benzoyl-L-arginine ethyl ester hydrolysis. These prototropic groups had pKa values of 4.8, 6.5 to 6.7, and 8.7. Theoretical treatment of the kinetics provided an excellent fit with the experimentally found profile when the contribution of all three prototropic groups was considered. Analysis showed that, in acid, the pathways of papain catalysis were functional with either two or three active-site protons. In base, a single functional ionic pathway is associated with an active site with only one proton. Pathways involving an unprotonated active site are catalytically inoperative in both acid and base. These results indicate that papain exhibits several catalytically functional ionic pathways. The results are discussed in terms of pKa assignments, and the mechanism of papain catalysis.  相似文献   

15.
The 11-cis-retinal binding site of rhodopsin is of great interest because it is buried in the membrane but yet must provide an environment for charged amino acids. In addition, the active-site lysine residue must be able to engage in rapid Schiff base formation with 11-cis-retinal at neutral and lower pH values. This requires that this lysine be unprotonated. We have begun to study the environment of the active-site lysine using a reporter group adducted to it. Non-active-site permethylated opsin was reacted with 5-nitrosalicylaldehyde, and the resulting Schiff base was permanently fixed by borohydride reduction. The stoichiometry of incorporation was one. This chromophoric and pH-sensitive reporter group affords information on the active-site environment of rhodopsin by determining the ionization constants of its ionizable groups at different pH values. The pH titration of the modified protein showed a single pKa = 7.8 +/- 0.19 ascribable to the ionization of the phenol. The ionization of the modified lysine residue was not observed at all pH values studied. These studies are interpreted to mean that a negatively charged amino acid is propinquous to the active-site lysine residue and that this latter residue does not have an unusually low pKa.  相似文献   

16.
The active site of glucosamine-6-phosphate deaminase (EC 3.5.99.6, formerly 5.3.1.10) from Escherichia coli was first characterized on the basis of the crystallographic structure of the enzyme bound to the competitive inhibitor 2-amino-2-deoxy-glucitol 6-phosphate. The structure corresponds to the R allosteric state of the enzyme; it shows the side-chain of His143 in close proximity to the O5 atom of the inhibitor. This arrangement suggests that His143 could have a role in the catalysis of the ring-opening step of glucosamine 6-phosphate whose alpha-anomer is the true substrate. The imidazole group of this active-site histidine contacts the carboxy groups from Glu148 and Asp141, via its Ndelta1 atom [Oliva et al. (1995) Structure 3, 1323-1332]. These interactions change in the T state because the side chain of Glu148 moves toward the allosteric site, leaving at the active site the dyad Asp141-His143 [Horjales et al. (1999) Structure 7, 527-536]. In this research, a dual approach using site-directed mutagenesis and controlled chemical modification of histidine residues has been used to investigate the role of the active-site histidine. Our results support a multifunctional role of His143; in the forward reaction, it is involved in the catalysis of the ring-opening step of the substrate, glucosamine 6-P. In the reverse reaction, the substrate fructose 6-P binds in its open chain, carbonylic form. The role of His143 in the binding of both glucosamine 6-P and reaction intermediates in their extended-chain forms was demonstrated by binding experiments using the reaction intermediate analogue, 2-amino-2-deoxy-D-glucitol 6-phosphate. His143 was also shown to be a critical residue for the conformational coupling between active and allosteric sites. From the pH dependence of the reactivity of the active site histidine to diethyl dicarbonate, we observed a pK(a) change of 1.2 units to the acid side when the enzyme undergoes the allosteric T to R transition during which the side chain of Glu148 moves toward the active site. The kinetic study of the Glu148-Gln mutant deaminase shows that the loss of the carboxy group and its replacement with the corresponding amide modifies the k(cat) versus pH profile of the enzyme, suggesting that the catalytic step requiring the participation of His143 has become rate-limiting. This, in turn, indicates that the interaction Glu148-His143 in the wild-type enzyme in the R state contributes to make the enzyme functional over a wide pH range.  相似文献   

17.
Human immunodeficiency virus type 1 (HIV-1) protease optimally catalyzes in the pH range of 4-6 in contrast to nearly all of the other eukaryotic aspartic proteases, which catalyze best in the pH range of 2-4. A possible structural reason for the higher optimal pH of HIV-1 protease is the absence of a hydrogen bond to the carboxyl group of active-site Asp25, which is nearly universally present in others. To investigate this hypothesis, we have mutated residue 28 in HIV-1 protease from alanine to serine. Both the wild-type and the mutant A28S enzymes have been overexpressed in Escherichia coli using a chemically synthesized gene and purified for a comparative study in enzyme kinetics. The kcat and Km values were determined by a radiometric assay for the wild-type enzyme from pH 3.2 to 7.0, and for the mutant enzyme from pH 3.2 to 6.0. The low pK values of the active site of the free enzyme, pKe1, are 3.3 and 3.4 for the wild-type and mutant enzymes, respectively. The low pK values of the active site of the enzyme bound to substrate, pKes1, are 5.1 and 4.3 for the wild-type and mutant enzymes, respectively. The high pK values of the free enzyme, pKe2, are 6.8 and 5.6, and the corresponding ones for the substrate-bound enzyme, pKes2, are 6.9 and 6.0 for the wild-type and mutant enzymes, respectively. The lowering of pK values in mutant HIV-1 protease indicates that the hydroxyl group of Ser28 forms a new hydrogen bond to active-site Asp25 to increase its acidity.  相似文献   

18.
The interaction of aromatic donor molecules with manganese(III) protoporphyrin-apohorseradish peroxidase complex [Mn(III)HRP] was investigated by optical difference spectroscopy and relaxation rate measurements of 1H resonances of aromatic donor molecules (at 500 MHz). pH dependence of substrate proton resonance line-widths indicated that the binding was facilitated by protonation of an amino acid residue (with a pKa of 6.1), which is presumably distal histidine. Dissociation constants were evaluated from both optical difference spectroscopy and 1H-NMR relaxation measurements (pH 6.1). The dissociation constants of aromatic donor molecules were not affected by the presence of excess of I-, CN- and SCN-. From competitive binding studies it was shown that all these aromatic donor molecules bind to Mn(III)HRP at the same site, which is different from the binding site of I-, CN- and SCN-. Comparison of the dissociation constants between the different substrates suggests that hydrogen bonding of the donors with distal histidyl amino acid and hydrophobic interaction between the donors and active site contribute significantly towards the associating forces. Free energy, entropy and enthalpy changes associated with the Mn(III)HRP-substrate equilibrium have been evaluated. These thermodynamic parameters were found to be all negative. Distances of the substrate protons from the paramagnetic manganese ion of Mn(III)HRP were found to be in the range of 7.7 to 9.4 A. The Kd values, the thermodynamic parameters and the distances of the bound aromatic donor protons from metal center in the case of Mn(III)HRP were found to be very similar as in the case of native Fe(III)HRP.  相似文献   

19.
A general route for the synthesis of chloromethyl ketone derivatives of fatty acids is described. 5-Chloro-4-oxopentanoic acid, 7-chloro-6-oxoheptanoic acid, 9-chloro-8-oxononanoic acid and 11-chloro-10-oxoundecanoic acid were synthesized by this method and tested as covalent inhibitors of pig heart acetoacetyl-CoA thiolase. The K1 decreased by approx. 20-fold for each pair of methylenes added to the chain length, showing that the initial stage in inhibitor binding occurs at a non-polar region of the protein. This region is probably located at the enzyme active site, since inhibition was prevented by acetoacetyl-CoA or acetyl-CoA but not by CoA. The site of modification by chloromethyl ketone derivatives of fatty acids is restricted to a thiol group, since inactivation of the enzyme was prevented by reversible thiomethylation of the active-site thiol. In contrast, an amino-directed reagent, citraconic anhydride, still inactivated the enzyme, even when the active-site thiol was protected. Evidence that the enzyme thiol was particularly reactive came from studies on the pH-dependence of the alkylation reaction and thiol-competition experiments. Inhibition of the enzyme proceeded suprisingly well at acidic pH values and a 10(5) molar excess of external thiol over active-site thiol was required to prevent inhibition by 0.3 mM-9-chloro-8-oxononanoic acid. In addition to inhibiting isolated acetoacetyl-CoA thiolase, in hepatocytes the chloromethyl ketone derivatives of fatty acids also inhibited chloresterol synthesis, which uses this enzyme as an early step in the biosynthetic pathway. In isolated cells, the chloromethyl ketone derivatives of fatty acids were considerably less specific in their inhibitory action compared with 3-acetylenic derivatives of fatty acids, which act as suicide inhibitors of acetoacetyl-CoA thiolase. However, 9-chloro-8-oxononanoic acid was also an effective inhibitor of both hepatic cholesterol and fatty acid synthesis in mice in vivo, whereas the acetylenic fatty acid derivative, dec-3-ynoic acid, was completely ineffective. The effective inhibitory dose of 9-chloro-8-oxononanoic acid (2.5-5 mg/kg) was substantially lower than the estimated LD50 for the inhibitor (100 mg/kg).  相似文献   

20.
Peptide deformylase (PDF) catalyzes the hydrolytic removal of the N-terminal formyl group from nascent ribosome-synthesized polypeptides in eubacteria. PDF represents a novel class of mononuclear iron protein, which utilizes an Fe(2+) ion to catalyze the hydrolysis of an amide bond. This Fe(2+) enzyme is, however, extremely labile, undergoing rapid inactivation upon exposure to molecular oxygen, and is spectroscopically silent. In this work, we have replaced the native Fe(2+) ion with the spectroscopically active Co(2+) ion through overexpression in the presence of Co(2+). Co(2+)-substituted PDF (Co-PDF) has an activity 3-10-fold lower than that of the Fe(2+)-PDF but is highly stable. Steady-state kinetic assays using a series of substrates of varying deformylation rates indicate that Co-PDF has the same substrate specificity as the native enzyme. Co-PDF and Fe-PDF also share the same three-dimensional structure, pH sensitivity, and inhibition pattern by various effector molecules. These results demonstrate that Co-PDF can be used as a stable surrogate of Fe-PDF for biochemical characterization and inhibitor screening. The electronic absorption properties of the Co(2+) ion were utilized as a probe to monitor changes in the enzyme active site as a result of site-directed mutations, inhibitor binding, and changes in pH. Mutation of Glu-133 to an alanine completely abolishes the catalytic activity, whereas mutation to an aspartate results in only approximately 10-fold reduction in activity. Analysis of their absorption spectra under various pH conditions reveals pK(a) values of 6.5 and 5.6 for the metal-bound water in E133A and E133D Co-PDF, respectively, suggesting that the metal ion alone is capable of ionizing the water molecule to generate the catalytic nucleophile, a metal-bound hydroxide. On the other hand, substrate binding to the E133A mutant induces little spectral change, indicating that in the E.S complex the formyl carbonyl oxygen is not coordinated with the metal ion. These results demonstrate that the function of the active-site metal is to activate the water molecule, whereas Glu-133 acts primarily as a general acid, donating a proton to the leaving amide ion during the decomposition of the tetrahedral intermediate.  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