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 968 毫秒
1.
Diamine oxidase from Euphorbia characias latex contains two sulfhydryl groups per mole of dimeric enzyme. The sulfhydryl groups are unreactive in the native enzyme but can be readily titrated by 4,4′-dithiodipyridine after protein denaturation, or anaerobically in the presence of the amine substrate. In the presence of both substrates (diamine and oxygen) they react sluggishly. The sulfhydryl groups show different reactivity toward various reagents, but in every case their modification inhibits catalytic activity. The insensitivity of the native enzyme to specific reagents suggests that the sulfhydryl groups are positioned in the interior of the protein and shielded from the solvent. Their reactivity in the presence of the amine substrate could be attributed to a conformational change occurring upon substrate binding or after substrate oxidation.  相似文献   

2.
The acetyl-CoA:acetoacetate CoA-transferase of Escherichia coli has the subunit structure α2β2 The enzyme contains six sulfhydryl groups, one per α chain and two per β chain, and no disulfides. The rates and extent of sulfhydryl group reactivity with 5,5′-dithiobis(2-nitrobenzoic acid) were compared in the free enzyme, the enzyme-CoA intermediate in the catalytic pathway, and a substrate analog-enzyme Michaelis complex. The analog used was acetylaminodesthio-CoA, a competitive inhibitor with respect to acetyl-CoA; the analog is not a substrate. The reactions were studied in the presence and absence of 10% glycerol. In the absence of glycerol, one sulfhydryl group reacted rapidly in the free enzyme and enzyme-CoA intermediate; relative to the free enzyme, the rate and number of subsequently reacting sulfhydryl groups were increased in the enzyme-CoA intermediate. In the presence of 10% glycerol, one sulfhydryl group reacted rapidly in the free enzyme, while two reacted rapidly in the enzyme-CoA compound; the rates and extents of subsequently reacting sulfhydryl groups were also enhanced in the enzyme-CoA compound. The data strongly suggested subunit interactions in the free enzyme and intermediate; glycerol abolished those interactions in the enzyme-CoA intermediate. In the absence of glycerol, sulfhydryl group reactivity in the Michaelis complex, enzyme-acetylaminodesthio-CoA, was similar to that in the free enzyme with one exception: One of the more slowly reacting sulfhydryl groups in the free enzyme reacted at a rate characteristic of the enzyme-CoA intermediate. The results obtained with N-ethylmaleimide were qualitatively similar. The fractional inactivation of the enzyme with N-ethylmaleimide as a function of sulfhydryl groups modified and the subunit location of those sulfhydryl groups indicated that the same sulfhydryl groups react in both enzyme species; however, those sulfhydryl groups reacted more rapidly in the enzyme-CoA compound. The data indicate both subunit interactions in the enzyme and characteristic conformational changes upon formation of an acyl-CoA-enzyme Michaelis complex and the enzyme-CoA intermediate.  相似文献   

3.
It is known that the enzymatic activity of papain (EC 3.4.22.2) toward α-N-benzoyl-l-arginine p-nitroanilide can be substantially increased by hydroxynitrobenzylation of Trp-177 through reaction of the enzyme with the active site-directed reagent, 2-chloromethyl-4-nitrophenyl (N-carbobenzoxy)glycinate (S.-M. T. Chang and H. R. Horton, 1979, Biochemistry18, 1559–1563). To determine the effect of such hydroxynitrobenzylation on the nucleophilicity of the essential thiol group at the active site of the enzyme, rates of inactivation by SN2 reactions of Cys-25 with chloroacetamide and chloroacetate and by Michael addition of Cys-25 to N-ethylmaleimide were monitored. The kinetics revealed that, at pH 6.5, the reactivities of the sulfhydryl group of hydroxynitrobenzylated papain with chloroacetamide and with N-ethylmaleimide are 24 and 27% greater than those of the sulfhydryl group of native papain. At pH 7.1, the rate enhancements are 34 and 39%, respectively. These increases in reactivity of Cys-25 as an attacking nucleophile appear to account for the increased catalytic activity of hydroxnitrobenzyl-papain toward an oligopeptide substrate, α-N-benzoyl-l-phenylalanyl-l-valyl-l-arginine p-nitroanilide, and toward an ester substrate, N-carbobenzoxyglycine p-nitrophenyl ester. However, the presence of the hydroxynitrobenzyl reporter group provides substantially greater improvement (250%) in enzymatic efficiency toward α-N-benzoyl-l-arginine p-nitroanilide, apparently by blocking nonproductive binding of this substrate to the enzyme. Fluorescence changes accompanying the various chemical modifications are interpreted in terms of a charge-transfer interaction between the imidazolium ion of His-159 and the indole moiety of Trp-177 in the active form of native papain, which should help to stabilize the catalytically essential mercaptide-imidazolium ion-pair (Cys-25, His-159).  相似文献   

4.
The acetyl-CoA:acetoacetate CoA-transferase of Escherichia coli undergoes two detectable conformational changes during catalysis of CoA transfer. The first change occurs upon binding of at least the CoA moiety of an acyl-CoA substrate and was detected by fluorescence enhancement of enzyme-bound 8-anilino-1-naphthalenesulfonate and microcomplement fixation upon formation of a noncovalent enzyme · CoA complex. CoA is a competitive inhibitor with respect to acyl-CoA substrate (Ki = 0.29 mM). A second, more extensive conformational change occurs upon formation of the covalent enzyme-CoA intermediate and was detected by fluorescence enhancement of enzymebound 8-anilino-1-naphthalenesulfonate, sedimentation of the intermediate in sucrose density gradients, and microcomplement fixation. The data clearly differentiated between the three distinct forms of the enzyme, i.e., free enzyme, noncovalent enzyme·CoA complex, and covalent enzyme-CoA intermediate. The data are consistent with a model in which the enzyme opens upon formation of the enzyme-CoA intermediate. Either the limited conformational change or the extensive conformational change generates subunit interactions which result in half-the-sites reactivity in the enzyme. Only one of the two potential active sites was charged with etheno-CoA when the enzyme was reacted with etheno-acetyl-CoA. Glycerol abolished the extreme negative cooperativity and both active sites were charged with etheno-CoA in the presence of 10% glycerol. Our data suggest that glycerol abolished subunit interactions in either the enzyme-CoA complex or the covalent intermediate and not in the free enzyme.  相似文献   

5.
The V of horse liver aldehyde dehydrogenase is enhanced twofold in the presence of 0.5 mm Mg2+ ions when assayed in the dehydrogenase reaction. The mechanism of this activation appears to be related to the fact the enzyme changes from functioning with half-of-the-sites reactivity to functioning with all-of-the-sites reactivity. That is, the presteady-state burst magnitude increases from 2 mol NADH formed per mole of tetrameric enzyme to 4 mol formed per mole (K. Takahashi and H. Weiner, J. Biol. Chem., 1980, 255, 8206–8209). Whether this twofold enhancement correlates, in fact, to a change from half-of-the-sites to all-of-the-sites reactivity of the enzyme by Mg2+ ions was investigated by determining the Stoichiometry of coenzyme binding by fluorescence quenching and enhancement methods in the absence and presence of the metal ions. The biphasic Scatchard plots for NAD binding to the enzyme were similar in the absence and presence of Mg2+ ions, while that of NADH binding was monophasic (-Mg2+) and biphasic (+Mg2+). In the presence of p-methoxyacetophenone, a competitive inhibitor for substrate, the stoichiometric titration of coenzyme binding to the ternary complexes (enzyme-NAD(H)-inhibitor) revealed that only 2 mol of NAD or NADH bind in the absence of Mg2+ ions but 4 bind per mole of tetrameric enzyme in the presence of added metal. The fluorescence intensity of NAD's fluorescent derivative, 1,N6-ethenoadenine dinucleotide, bound to the enzyme was also doubled by the addition of Mg2+ ions.The combined binding data show that the stoichiometry of coenzyme binding to aldehyde dehydrogenase in the ternary complex increases from 2 to 4 mol binding per mole of tetrameric enzyme with the addition of Mg2+ ions. This increase in stoichiometry corresponds to the observed changes of burst magnitude obtained from the presteady-state and V in the steady-state kinetics assays. From both results of the kinetics and stoichiometry, we show that horse liver aldehyde dehydrogenase exhibits half-of-the-sites reactivity when in the tetrameric state in the absence of Mg2+ ions, and all-of-the-sites reactivity in the dimeric state in the presence of the metal.  相似文献   

6.
The free catalytic subunit of cAMP-dependent protein kinase readily undergoes a pronounced, salt-induced conformational change at neutral pH and around physiological values of ionic strength. This change, which is fully reversible, can be monitored directly by the relative chemical reactivity of two SH groups in the enzyme. Upon increasing the ionic strength of the medium from 0.03 to 0.22, one sulfhydryl becomes more reactive towards 5,5′-dithiobis[2-nitrobenzoic acid] while the other sulfhydryl becomes less reactive towards the same reagent. In parallel, the enzyme undergoes a salt-induced inactivation when histone H2b is used as a substrate. Though not reflected in the Vmax, this conformational change considerably increases the Km of the enzyme for histone H2b as well as for MgATP. This intrinsic malleability of the enzyme can account for the well-known salt inhibition of the enzyme for certain substrates and ion-dependent activation towards other substrates. It is suggested that this malleability might constitute the molecular basis for modulating the specificity of the enzyme and channeling its activity from one substrate to another in response to intracellular specifier signals.  相似文献   

7.
One sulfhydryl group of the mitochondrial isoenzyme of aspartate aminotransferase from both chicken and pig heart exhibits syncatalytic reactivity changes similar to those found previously in the cytosolic isoenzyme from pig heart (Birchmeier, W., Wilson, K.J., and Christen, P. (1973) J. Biol. Chem. 248, 1751–1759). The reactivity of the only titratable sulfhydryl group toward 5,5′-dithiobis-(2-nitrobenzoate) is at a minimum in the free pyridoxal and pyridoxamine form of the enzyme and is increased by approximately one order of magnitude when covalent enzyme-substrate intermediates are formed. The modification of the sulfhydryl group does not affect enzymatic activity. This finding supports the earlier conclusion that the syncatalytic reactivity changes are not due to a direct participation of this group in the active site but rather to conformational adaptations of the enzyme-coenzyme-substrate compound occurring in the catalytic mechanism of aspartate aminotransferases.  相似文献   

8.
Orthorhombic single crystals of cytoplasmic aspartate aminotransferase were examined alone or in the presence of substrates or inhibitors to quantitatively compare the interaction of ligands with the active-site chromophore between soluble and crystalline enzyme. As in enzyme solutions, equilibrium kinetic measurements can be made between substrates and single crystals of cytoplasmic aspartate aminotransferase. The absorption spectra of ligand-free enzyme forms and of enzyme-substrate or-inhibitor complexes are as distinctive as when the enzyme is in solution. The dissociation constants for glutamate with the pyridoxal form of the enzyme are identical to those in solution. The substrate analog erythro--hydroxyaspartate also binds with equal affinity to the active site in enzyme crystals as in solution; and the affinity of -ketoglutarate to bind in nonproductive complexes with the pyridoxal form of the enzyme is also unimpaired in the crystal (K d =2 mM). In contrast to the affinity constants, the stoichiometry of the interactions does not appear to correlate to those in solution. In the presence of an amino acid plus keto acid substrates pair, the absorbance values of the enzyme-substrate complex(es) could be interpreted as for occupany of only half the available sites in the crystals. Yet an amino acid, cysteine sulfinate, and -keto acids such as , -difluorooxalacetate convert all active sites in the crystal to the pyridoxamine or pyridoxal form when added to the pyridoxal or pyridoxamine forms, respectively. This ability to completely undergo substrate-induced half-transamination and the apparently conflicting results in trapping half the sites in enzyme-substrate complexes are incorporated into a proposed reciprocating mechanism applicable only to the crystalline state of the enzyme and dictated by crystal packing forces rather than an intrinsic property of the enzyme. Active-site bound pyridoxal phosphate continues to behave as a pH indicator; nevertheless, the pK value of the single crystals is a pH unit (pK=7.15) higher than that in solution. This variation is interpreted as indication of a difference in the environment of the chromophore between the crystal and solution states. While the environmental difference does not significantly alter the affinity for substrates, it could account for the reduced rates in transformation of the enzyme-substrate complexes in half-transamination reactions in the crystalline state.  相似文献   

9.
Rat liver tyrosine aminotransferase and alanine aminotransferase are similar enzymes in most properties, but they differ markedly in their ease of coenzyme dissociation and rate of metabolic turnover. Dissociation of coenzyme does not determine rate of turnover (K.L. Lee, P. L. Darke, and F. T. Kenney, 1977, J. Biol. Chem.252, 4958–4961), but these parameters may reflect structural properties of the enzymes which determine both. To explore this possibility we studied these enzymes in livers of rats fed a pyridoxine-deficient diet in which both enzymes were largely in apoenzyme form. This form of alanine aminotransferase, not previously characterized, was identified as an immunologically cross-reactive material which was converted to active enzyme when extracts were incubated with pyridoxal phosphate in vitro. This apoenzyme behaved like the active holoenzyme in chromatographic and electrophoretic analyses but was more sensitive than the holoenzyme to heat, low pH, or proteolysis by trypsin or chymotrypsin. Relative rates of reconstitution of the two holoenzymes in vivo after injection of pyridoxine were determined as a measure of conformational stability of the two enzymes as they exist in the intracellular environment. Restoration of the tyrosine aminotransferase holoenzyme was completed within 30 to 45 min, but that of the alanine enzyme required 8 h. These results suggest that tyrosine aminotransferase in vivo is a relaxed structure which facilitates both coenzyme dissociation and rapid metabolic turnover, whereas alanine aminotransferase assumes a taut structure resistant to both dissociation and degradative processes.  相似文献   

10.
The kinetic theory of the substrate reaction during modification of enzyme activity previously described by Tsou [Tsou (1988),Adv. Enzymol. Relat. Areas Mol. Biol. 61, 381–436] has been applied to a study of the kinetics of the course of inactivation of the mitochondrial succinate-ubiquinone reductase by 5,5′-dithiobis-(2-nitro-benzoic acid) (DTNB). The results show that the inactivation of this enzyme by DTNB is a conformation-change-type inhibition which involves a conformational change of the enzyme before inactivation. The microscopic rate constants were determined for the reaction of the inactivator with the enzyme. The presence of the substrate provides marked protection of this enzyme against inactivation by DTNB. The modification reaction of the enzyme using DTNB was shown to follow a triphasic course by following the absorption at 412 nm. Among these reactive thiol groups, the fast-reaction thiol group is essential for the enzyme activity. The results suggest that the essential thiol group is situated at the succinate-binding site of the mitochondrial succinate-ubiquinone reductase.  相似文献   

11.
The 16 sulfhydryl groups of native, homogeneous rabbit muscle fructose diphosphatase can all react with 5,5′-dithiobis-(2-nitrobenzoic acid). High concentrations of substrate (1–2 mm) decrease the reaction rate of the sulfhydryl groups, while concentrations up to 70 μm have no effect. After titration of the four most rapidly reacting sulfhydryl groups there is a marked desensitization toward the allosteric inhibitor AMP. In the presence of 30 μm AMP only 4–5 sulfhydryl groups/tetramer react with 5,5′-dithiobis-(2-nitrobenzoic acid), and the enzyme again becomes desensitized toward AMP inhibition. Together with a 3.5-fold increase in the I50 for AMP inhibition, the Km for Mg2+ or Mn2+ ions is also increased. In the presence of 7 mm MgCl2 or 0.28 mm MnCl2 only 6–8 sulfhydryl groups are modified. The rapid reaction of 4 sulfhydryl groups again results in desensitization. There is neither a protection by the substrate against inactivation, nor a protection by the allosteric inhibitor against desensitization. It is concluded that AMP and the divalent cations induce conformational changes in the protein molecule making 11–12 or 8–10 sulfhydryl groups inert for 5,5′-dithiobis-(2-nitrobenzoic acid), respectively. The Km for fructose-1,6-diphosphate is not changed after the modification of 4–5 sulfhydryl groups.  相似文献   

12.
Dihydrofolate reductase from chicken liver has a single sulfhydryl group which reacts stoichiometrically and specifically with a wide variety of organic mercury compounds to yield an enzyme derivative which exhibits up to 10-fold the activity of the unmodified form when measured at pH 6.5, the optimum for the modified enzyme. The sulfhydryl group is apparently not at the active site since a 25-fold excess of either major cosubstrate, dihydrofolate or TPNH, affects neither the rate nor extent of the modification reaction. The reaction is essentially instantaneous and yields an enzyme with altered kinetic properties for all the substrate pairs examined (TPNH/dihydrofolate, TPNH/ folate, and DPNH/dihydrofolate) when tested near their pH optima. V values increased 3- to 10-fold when TPNH was cofactor; Km values increased 10- to 15-fold for the TPNH/dihydrofolate pair. The mercurial-activated enzyme, unlike the native form, exhibits a markedly increased sensitivity to heat, proteolysis, and the ionic environment, losing approximately 50% of its activity under conditions where there is no loss of activity in the native form. However, substrates can afford protection, the order of effectiveness being identical with the relative affinities of the substrates for the native enzyme (Subramanian, S., and Kaufman, B. T. (1978) Proc. Nat. Acad. Sci. USA75, 3201). Thus, dihydrofolate, with the largest binding constant is the most efficient, protecting completely against trypsin digestion when present at a 1:1 ratio with enzyme. Heating the mercury enzyme in the absence of substrates gives rise to a stable but altered conformation characterized by a time course which shows marked hysteresis. The striking similarity of the properties of the mercurial-activated dihydrofolate reductase to the reductase activated by 4 m urea, a reagent known to affect the tertiary structure of proteins, suggests that covalent binding of organic mercurials to the sulfhydryl group results in a similar conformational change characterized by a marked facilitation of the dihydrofolate reductase reaction.  相似文献   

13.
A study of the sulfhydryl groups of rat brain hexokinase   总被引:1,自引:0,他引:1  
Rat brain hexokinase (ATP: d-hexose-6-phosphotransferase, EC 2.7.1.1) is rapidly inactivated by reaction with 5,5′-dithiobis-(2-nitrobenzoate). The inactivation follows monophasic first-order kinetics in either the absence of ligands (k = 0.641 min?1 at 25 °C) or in the presence of saturating levels of ATP (free or complexed with Mg2+) or P1; the inactivation rate is slightly increased (k ? 0.7 min ?1) in the presence of ATP or P1. In contrast, glucose and glucose-6-P markedly decrease the inactivation rate; inactivation in the presence of these ligands is biphasic, with two first-order rates (k ? 0.5 min?1 and 0.01 min?1) being distinguishable.The enzyme contains 14 sulfhydryl groups which react with 5,5′-dithiobis-(2-nitrobenzoate); reaction of these groups in the native enzyme is complete after 2 hr at 25 °C, or in approx 5 min with the urea or guanidine-denatured enzyme. In the native enzyme, three classes of sulfhydryl groups are distinguishable and are designated as F-, I-, or S-type based on their fast (k ? 0.7 min?1), intermediate (k ? 0.5-0.7 min?1), or slow (k ? 0.02 min?1 rates of reaction with 5,5′-dithiobis-(2-nitrobenzoate). The correlation of inactivation rates with the rates for reaction of the I-type sulfhydryls indicates that the I-type sulfhydryls include residues necessary for catalytic activity. The F-type residues are clearly not required for activity.The effects of ATP, P1, glucose, and glucose-6-P on the reactivity of the sulfhydryls have been determined. As in the absence of ligands, S-, I-, and F-type sulfhydryls could be distinguished. In the presence of saturating concentrations of these ligands, the F, I, and S classes of sulfhydryls contained respectively: with ATP, 1, 4, and 7 residues; with P1, 1, 3, and 7 residues; with glucose, 1, 2, and 5 residues; with glucose-6-P, 1, 2, and 1 residues. Comparison with rate constants for inactivation in the presence of these ligands again indicated that I-type sulfhydryls were particularly important in maintenance of enzyme activity. The present results indicate considerable similarity between the reactivity of the sulfhydryl residues in rat brain hexokinase and the sulfhydryls of the bovine brain enzyme [V. D. Redkar and U. W. Kenkare (1972), J. Biol. Chem., 247, 7576–7584].  相似文献   

14.
Inactivation of rat brain hexokinase (ATP:d-hexose 6-phosphotransferase, EC 2.7.1.1) by the arginine-specific reagent, phenylglyoxal, has been studied. Inactivation did not follow pseudo-first-order kinetics, suggesting the involvement of two or more arginine residues in catalytic function. Using [14C]phenylglyoxal, it was found that 5 of the 55 arginines per molecule of hexokinase react with this reagent, with an accompanying loss of over 90% of the catalytic activity. Virtually all of the activity loss occurs during derivatization of four relatively slower reacting arginines, with essentially no activity loss during derivatization of one rapidly reacting arginine. Inactivation by phenylglyoxal was not due to reaction with critical sulfhydryl groups in brain hexokinase since reactivity of the enzyme with the sulfhydryl reagent, 5,5′-dithiobis(2-nitrobenzoic acid) was not affected by prior treatment with phenylglyoxal. Comparison of amino acid composition, before and after reaction with phenylglyoxal, indicated that only the arginine content had been affected by phenylglyoxal treatment. The decrease in arginine content, measured by amino acid analysis, and the incorporation of phenylglyoxal, measured with [14C]phenylglyoxal, was consistent with the phenylglyoxal:arginine stoichiometry of 2:1 originally reported by K. Takahashi (1968, J. Biol. Chem.243, 6171–6179). Several ligands were tested and found to provide varying degrees of protection of hexokinase activity against phenylglyoxal. ATP and ADP alone provided only slight protection, but were highly effective in the presence of N-acetylglucosamine which itself gave only moderate protection. Glucose 6-phosphate and 1,5-anhydroglucitol 6-phosphate, both good inhibitors of brain hexokinase, were very effective while poorly inhibitory hexose 6-phosphates were not. Glucose was very effective, with protection afforded by other hexoses being correlated with their ability to serve as substrates (i.e., poor substrates also provided little protection against phenylglyoxal). The effectiveness of hexose 6-phosphates and hexoses in protecting the enzyme against inactivation by phenylglyoxal was related to their ability to induce conformational change in the enzyme. None of the ligands tested appreciably affected the reactivity of the rapidly reacting arginine residue. There was no correlation between the inhibition observed in the presence of various ligands and the number of arginines reacted with phenylglyoxal. The results were interpreted as indicating the involvement of two to four arginine residues in the catalytic function of brain hexokinase, possibly in the binding of anionic ligands such as ATP, ADP, or glucose 6-phosphate.  相似文献   

15.
Two of the 5 sulfhydryl residues of the β2 subunit of tryptophan synthase have previously been shown to react with N-ethylmaleimide and to have active site roles. We now show that the single sulfhydryl which reacts with N-ethylmaleimide in the presence of pyridoxal phosphate is cysteine-170. The essential sulfhydryl which reacts with N-ethylmaleimide or with 2-nitro-5-thiocyanobenzoic acid after removal of pyridoxal phosphate is cysteine-230. The affinity reagent, bromoacetylpyridoxamine phosphate, reacts variably with cysteine-62 or with cysteine-230.  相似文献   

16.
Sulfhydryl oxidase (glutathione-oxidizing activity) is closely associated with γ-glutamyltransferase (γ-glutamyl transpeptidase) in skim milk membranes. Similar close association of the two enzymatic activities in kidney membranes has led to the recent proposal that glutathione-oxidizing activity can be attributed to the action of γ-glutamyltransferase, itself, in generating cysteinylglycine which, in turn, catalyzes sulfhydryl group oxidation (O. W. Griffith and S. S. Tate, 1980, J. Biol. Chem.255, 5011–5014). However, a previously published procedure for the isolation of highly purified sulfhydryl oxidase from skim milk membranes (V. G. Janolino and H. E. Swaisgood, 1975, J. Biol. Chem.250, 2532–2538) leads to the effective separation of the two activities. Quantitative chromatographic analyses of GSH, GSSG, and Glu levels revealed that the highly purified sulfhydryl oxidase preparation catalyzes the direct oxidation of GSH to GSSG without detectable cleavage of the γ-glutamyl peptide bond. These results were confirmed by monitoring the time course of substrate disappearance and product formation using high-performance liquid chromatography. Conversely, a supernatant fraction enriched in γ-glutamyltransferase activity displayed no sulfhydryl group-oxidizing activity. 6-Diazo-5-oxo-l-norleucine selectively inhibited the transferase in crude preparations containing both sulfhydryl oxidase and γ-glutamyltransferase. It is concluded that sulfhydryl oxidase and γ-glutamyltransferase activities are distinct and separable.  相似文献   

17.
The incubation of yeast alcohol dehydrogenase with formaldehyde in the presence of NaBH4 methylates lysine residues to form ?N,?N-dimethyl lysine with a concurrent decrease in enzymic activity which is not alleviated by the presence of coenzymes. The modification causes structural change(s) in yeast alcohol dehydrogenase as evidenced by a hyperchromic shift in the uv spectrum, the sensivitity to heat inactivation, the reactivity to sulfhydryl reagents, and a change in Stokes' radius. Kinetic studies indicate that the reduced activity of the methylated enzyme to oxidize alcohols is associated with decreased maximum velocities by retarding the interconversion of the ternary complexes. The catalytic efficiency of the control enzyme to oxidize primary alcohols is affected by the steric interaction which is absent in the methylated enzyme.  相似文献   

18.
The copper- and zinc-containing superoxide dismutase of bovine erythrocytes retains its native molecular weight of 32 000 in 8.0 M urea for at least 72 h at 25 degrees C, as evidenced by sedimentation equilibrium analysis. Subsequent to prolonged exposure to urea, the dimeric enzyme could be dissociated by sodium dodecyl sulfate in the absence of reductants, indicating the absence of unnatural disulfide cross-links. The sulfhydryl group of cysteine-6 was unreactive toward 5,5'-dithiobis(2-nitrobenzoic acid) or bromoacetic acid in both neutral buffer and 8.0 M urea. The histidine residues of the enzyme were resistant to carboxymethylation in neutral buffer and 8.0 M urea. However, when the enzyme was exposed to bromoacetic acid in the presence of 6.0 M guanidinium chloride and 1 mM (ethylenedinitriol)tetraacetic acid (EDTA), both sulfhydryl and histidine alkylation were observed. Guanidinium chloride (6.0 M) increased the reactivity of the sulfhydryl group of cysteine-6 and allowed the oxidative formation of disulfide-bridged dimers. This was prevented by 1 mM EDTA. It follows that 8.0 M urea neither dissociates the native enzyme into subunits nor produces a conformation detectably different than that possessed under native conditions.  相似文献   

19.
Racemic arylpropionic esters 13, precursors of therapeutically important non-steroidal antiinflammatory drugs, were subjected to hydrolyses in the presence of either Candida rugosa or Rhizomucor miehei crude lipases. The hydrolyses of 1 and 2 proved to be highly enantioselective, whereas 3 was not transformed at all. Both the substrate specificity and the enantioselectivity of these lipases were explained through a molecular modeling study involving docking experiments between 13 and the amino acids forming the enzymes active-sites, whose three-dimensional structures were obtained from X-ray crystallographic data, followed by extensive conformational analysis on their computer-generated complexes. The results of this study also account for the high enantioselective and good yielding hydrolysis of 3 (as the corresponding 2-chloroethyl ester) catalyzed by CRL pretreated with 2-propanol, recently reported in the literature, and lead to admit that such a treatment may operate very deep conformational changes on the amino acids of the enzyme active-site.  相似文献   

20.
When mitochondria are incubated with radioactively labeled mitochondrial aspartate aminotransferase (EC 2.6.1.1), the enzyme is taken up into the organelles. Mersalyl and p-hydroxymercuriphenyl sulfonic acid, but not N-ethylmaleimide or ethacrynic acid, decrease the extent of this uptake. Inhibition of the uptake by low concentrations of mercurial reagents is due to blockage of a single sulfhydryl group per monomer of the enzyme. Blockage of mitochondrial thiols does not inhibit uptake of the enzyme. A single sulfhydryl group out of a total of six per monomer of the native enzyme reacts with 5,5′-dithiobis-(2-nitrobenzoic acid). This is the same sulfhydryl group that reacts with low levels of mercurial reagents with consequent inhibition of uptake of the enzyme into mitochondria but without effect on the catalytic activity. N-Ethylmaleimide does not react with this group. N-Ethylmaleimide reacts with a different sulfhydryl group with concomitant decrease in enzymic activity but with no effect on uptake of the enzyme into mitochondria. High levels of mercurial reagents similarly decrease enzymic activity. Unlike the effect on uptake into mitochondria, the inhibition by mercurial reagents of enzymic activity is not reversed by treatment with cysteine. The significance of these observations with respect to the mechanism of uptake of aspartate aminotransferase into mitochondria is discussed, and comparisons are made between the reactivities of sulfhydryl groups in rat liver aspartate aminotransferase and in the enzymes from other animals.  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