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 93 毫秒
1.
Some preparations of both native aspartate transcarbamylase from Escherichia coli and catalytic subunit have fewer tight binding sites per oligomer for carbamyl-P than the number of catalytic peptide chains. In contrast, the number of sites for the tight-binding inhibitor N-(phosphonacetyl)-L-aspartate does equal the number of catalytic chains in each case. Binding of the labile carbamyl-P was determined using rapid gel filtration, with conversion to stable carbamyl-L-aspartate during collection. Native enzyme (six catalytic chains) obtained from cells grown under the conditions of J.C. Gerhart and H. Holoubek (J. Biol. Chem. (1967) 242, 2886-2892) has 5.4 tight sites for carbamyl-P at pH 8.0 (KD = 9.9 muM), whereas native enzyme from cells grown with higher concentrations of glucose, uracil, and histidine (to yield more enzyme per unit volume of culture) has only 1.9 tight sites at pH 8.0 (KD = 4.6 muM) and only 2.3 tight sites at pH 7.0 (KD = 2.6 muM). At pH 8.0, catalytic subunit (three catalytic chains) obtained from the former native enzyme has 2.2 tight sites for carbamyl-P (KD = 2.4 muM) and the number of sites is 2.3 in the presence of 35 mM succinate, whereas catalytic subunit obtained from the latter native enzyme has 1.8 tight sites (KD = 3.6 muM) in the absence of succinate and 2.3 tight sites in its presence. The number of tight binding sites is also less than the number of subunit peptide chains in 19F nuclear magnetic resonance experiments performed with catalytic subunit and two fluorinated analogs of carbamyl-P at comparable concentrations of analogs and active sites. A model is proposed in which incomplete removal of formylmethionine from the NH2 termini of the enzyme under conditions of extreme depression affects affinity for ligands.  相似文献   

2.
Nuclear magnetic relaxation methods were used to investigate the interaction of the inhibitor succinate with aspartate transcarbamylase from Escherichia coli. Over the pH range 7 to 9, the dissociation constant for succinate remains less than the inhibitor concentration used for most of this work (0.05 M). As a result, the enzyme predominantly exists in a single "gross" conformational state. Succinate binding to this enzyme state (generally known as the R form) parallels the behavior seen previously with the isolated catalytic subunit (Beard, C. B., and Schmidt, P.G. (1973) Biochemistry 12, 2255-2264). The pH and temperature dependence of succinate proton relaxation rates, 1/T2 - 1/T1, in the presence of carbamyl phosphate, is interpreted in terms of a binding mechanism involving three forms of the enzyme, differing by their states of protonation. The least protonated form of the enzyme does not interact with succinate, the singly protonated species binds succinate to form a rapidly dissociating complex, and the doubly protonated species undergoes a conformational isomerization upon succinate binding, yielding a slow exchange complex. Relaxation data provide sufficient information to determine pKa values of 7.2 and 8.9 for two ionizing groups, as well as the dissociation constant for succinate in the fast exchange complex, Kd =1.6 X 10(-2) M. Rate constants for the forward and reverse steps of the isomerization, 1.3 X 10(3) s-1 and 33 s-1, respectively, indicate a significantly slower reverse rate from that obtained in the earlier NMR study of the isolated catalytic subunit. In experiments where the succinate concentration was varied, the relaxation rates showed sigmoidal binding of that ligand to the fast exchange complex above pH 9.1, (a) indicating cooperative binding of succinate, and (b) suggesting that above pH 9.1, the system cannot be characterized by a single dissociation constant, ionization constant, or relaxation effect. CTP and ATP were tested for their ability to affect succinate binding to the fast exchange complex. Heterotropic interactions were observed for CTP but not for ATP. Addition of low concentrations of the transition state analog N-(phosphonacetyl)-L-aspartate to the enzyme-carbamyl phosphate-succinate complex sharply decreased the relaxation rate, indicating that the measurements are sensitive only to succinate bound specifically to the active site.  相似文献   

3.
Reaction of phenylglyoxal with aspartate transcarbamylase and its isolated catalytic subunit results in complete loss of enzymatic activity. This modification reaction is markedly influenced by pH and is partially reversible upon dialysis. Carbamyl phosphate or carbamyl phosphate with succinate partially protect the catalytic subunit and the native enzyme from inactivation by phenylglyoxal. In the native enzyme complete protection from inactivation is afforded by N-(phosphonacetyl)-L-aspartate. The decrease in enzymatic activity correlates with the modification of 6 arginine residues on each aspartate transcarbamylase molecule, i.e. 1 arginine per catalytic site. The data suggest that the essential arginine is involved in the binding of carbamyl phosphate to the enzyme. Reaction of the single thiol on the catalytic chain with 2-chloromercuri-4-nitrophenol does not prevent subsequent reaction with phenylglyoxal. If N-(phosphonacetyl)-L-aspartate is used to protect the active site we find that phenylglyoxal also causes the loss of activation of ATP and inhibition by CTP. The rate of loss of heterotropic effects is exactly the same for both nucleotides indicating that the two opposite regulatory effects originate at the same location on the enzyme, or are transmitted by the same mechanism between the subunits, or both.  相似文献   

4.
The solution properties of fibrinogen and the thrombin-induced activation and gelation of fibrinogen in 95% D2O at pH 7.4 were compared to those in H2O under similar conditions. The initial release rates of fibrinopeptides A and B in D2O were slightly slower than those in H2O. However, the values of the Michaelis-Menten parameters Km and V for the release of the two peptides in D2O and H2O in the presence of 0.5 M NaCl were about the same. From turbidity measurements at 450 nm it is obvious that fibrinogen is soluble in a slightly more narrow range of NaCl concentration and that the fibrin gels have a higher degree of lateral aggregation in D2O than in H2O. The variation of fibrinogen concentration, thrombin concentration, pH and ionic a strength have a similar dependence on the final gel structure and clotting time in D2O and H2O. SDS-gel electrophoresis on fibrin samples, which were cross-linked by factor XIII, yielded results where the cross-linking of the gamma-chain appeared to be the same in D2O and H2O. The alpha-chain cross-linking was somewhat faster in D2O than in H2O. When fibrinogen solutions in 95% D2O were incubated at 20 mM CaCl2, a slow gelation of fibrinogen was observed, which was found to be induced by trace amounts of factor XIII. The final gel turbidity appeared to be about the same for this gelation as for that induced by thrombin. The differences in solubility for fibrinogen, kinetics for the enzyme reaction and optical properties for the fibrin gels in D2O and H2O may be explained by differences in electrostatic interactions, hydrogen bonding and hydration of fibrinogen in these two media.  相似文献   

5.
13C and 15N isotope effects have been measured for the aspartate transcarbamylase (ATCase) reaction in an effort to elucidate the chemical mechanism of this highly regulated enzyme. The observed 15(V/K(asp))H2O value for the ATCase holoenzyme at saturing carbamyl phosphate and 12 mM L-aspartate is 1.0045 at pH 7.5, and this value remains unchanged in the presence of 5 mM ATP (activator) or 5 mM CTP (inhibitor). The fact that the isotope effect is not changed by the allosteric modifiers supports the conclusion that the kinetic properties of the active form of ATCase are not influenced by ATP or CTP. The observed 15(V/K(asp)) values for the catalytic subunit of ATCase are also the same as those determined for the holoenzyme, suggesting that the chemical mechanisms of both enzyme species are the same. Quantitative analysis of 13C and 15N isotope effects in both H2O and D2O has led to the proposal of a chemical model for the ATCase reaction which involves a precatalytic conformational change to form an activated complex that facilitates deprotonation of L-aspartate by an enzyme functional group. Nucleophilic attack on the carbonyl carbon of carbamyl phosphate by the alpha-amino group of L-aspartate results in the formation of a tetrahedral intermediate. An intramolecular proton transfer leads to formation of products N-carbamyl-L-aspartate and inorganic phosphate.  相似文献   

6.
The reaction of hydrogen peroxide with the copper-zinc bovine-liver superoxide dismutase at low molar ratios (0.2-20.0) of H2O2/active site between pH 7.3-10.0 leads to the loss of native enzyme as a distinct form monitored by electrophoresis. The pH dependence of the loss of native enzyme between 7.3 and 9.0 indicates the involvement of a conjugate base on the enzyme of pKa of 8.7 +/- 0.1. The rate of loss of the native enzyme is first order with respect to the concentration of both enzyme and hydrogen peroxide between pH 7.3 and 9.0 with no evidence for binding of peroxide. A second-order rate constant of 3.0 +/- 1.0 M-1 s-1 is obtained from these data. At pH 10.0 the reaction is first order with respect to enzyme concentration but saturable in H2O2. All data are consistent with the interpretation that H2O2 reacts with the enzyme at the lower pH where the reaction is dependent upon the conjugate base of a functional group on the enzyme. At the higher pH, the data are consistent with the reaction of HO2- and H2O2 with the dismutase. The dissociation constant for HO2- calculated from the kinetic data at pH 10.0 is between 25-50 microM and the rate constant for the breakdown of the HO2- dismutase complex is 1.10 + 0.05 x 10(-2) s-1. The change in the electrophoretic pattern at all pH values is accompanied by the loss of the ability of the enzyme to bind copper. Weakly bound or free copper can be detected using bathocuproine disulfonate. Furthermore copper-defficient forms of the enzyme can be detected by staining gels of the peroxide-treated dismutase with diethyldithiocarbamate.  相似文献   

7.
Aspartate transcarbamylase from Bacillus subtilis has been purified to apparent homogeneity. A subunit molecular weight of 33,500 +/- 1,000 was obtained from electrophoresis in polyarcylamide gels containing sodium dodecyl sulfate and from sedimentation equilibrium analysis of the protein dissolved in 6 M guanidine hydrochloride. The molecular weight of the native enzyme was determined to be 102,000 +/- 2,000 by sedimentation velocity and sedimentation equilibrium analysis. Aspartate transcarbamylase thus appears to be a trimeric protein; cross-linking with dimethyl suberimidate and electrophoretic analysis confirmed this structure. B. subtilis aspartate transcarbamylase has an amino acid composition quite similar to that of the catalytic subunit from Escherichia coli aspartate transcarbamylase; only the content of four amino acids is substantially different. The denaturated enzyme has one free sulfhydryl group. Aspartate transcarbamylase exhibited Michaelis-Menten kinetics and was neither inhibited nor activated by nucleotides. Several anions stimulated activity 2- to 5-fold. Immunochemical studies indicated very little similarity between B. subtilis and E. coli aspartate transcarbamylase or E. coli aspartate transcarbamylase catalytic subunit.  相似文献   

8.
The complex formed when excess regulatory subunits (r2) of aspartate transcarbamylase is added to a dilute solution of the catalytic subunit (c3) has been further studied. By stabilizing the complex with saturating levels or r2, it was possible to perform ultracentrifugation in sucrose density gradients. The sedimentation coefficient of the complex (7.7 plus or minus 0.2 S) is intermediate between those of the catalytic subunit (5.8 S) and of the native enzyme (11.7 S). Consideration of the likely hydrodynamic properties of the complex suggests that this sedimentation coefficient may be consistent with the c3r6 structure previously proposed. The formation of c3r6 from c3 and r2 is readily reversible. At nonsaturating levels or r2, conversion to the native enzyme (c3r6) takes place. This conversion is inhibited by high concentrations of r2. The c3r6 complex shows Michaelis-Menten kinetics with a low Km for aspartate and considerable substrate inhibition. The pH activity profile at high aspartate concentrations is almost identical with that of the native enzyme. All of these observations suggest that c3r6 represents the relaxed (R) state of aspartate transcarbamylase. The insensitivity of c3r6 toward CTP or ATP can also be explained by considering c3r6 as a stabilized relaxed state. These properties of c3r6 have significant implications regarding the allosteric mechanism of the native enzyme.  相似文献   

9.
Pyridoxal-P reacts specifically with a single lysine residue at the active site of Escherichia coli aspartate transcarbamylase (Greenwell, P., Jewett, S. L., and Stark, G. R. (1973) J. Biol. Chem. 248, 5994-6001). Reduction of the Schiff base with sodium borohydride, succinylation of the remaining lysine residues, and digestion with trypsin result in formation of a single pyridoxyl peptide, which was purified to homogeneity after chromatography on DEAE-cellulose, treatment with alkaline phosphatase, and rechromatography. Amino acid composition and the results of limited sequential degradation showed that this peptide corresponds to residues 62 to 98 in the sequence of Konigsberg and co-workers, and contains 2 residues of lysine (Henderson, L., Roy, D., Martin, D., and Konigsberg, W., personal communication). By similar isolation, a second peptide was obtained from unsuccinylated catalytic subunit, containing only the pyridoxylated lysine, which corresponds to Lys-80. Derivatives of catalytic subunit containing an average of either one, two, or three pyridoxamine-P moieties per trimer have been prepared by reduction. These species, which retain catalytic activity in proportion to their unmodified active sites, were recombined with regulatory subunit to prepare partially modified derivatives of native aspartate transcarbamylase. At pH 8, fluorescence emission bands were observed at 340 nm, due to aromatic amino acids in the protein, and at 395 nm, due to the pyridoxamine-P moiety. Upon excitation at 280 nm energy transfer from protein to pyridoxamine-P was approximately 15%. The properties of the probe were used to study changes accompanying the binding of substrates and inhibitors. The effects of CTP and ATP were small. With the transition state analog N-(phosphonacetyl)-L-aspartate (PALA) or the substrate carbamyl-P, two types of response were observed. Derivatives of catalytic subunit and native enzyme which contain some unmodified sites and hence retain partial catalytic activity gave large increases in fluorescence at 395 nm. However, fully modified inactive derivatives gave much smaller increases. A derivative of native enzyme containing one triply modified and one unmodified catalytic subunit behaved like the other partially modified species. These results indicate that there is communication among the active sites of different catalytic trimers in modified native enzyme, as well as among active sites within the same modified catalytic trimer. The increases in fluorescence result from a red shift of the absorption maximum of the pyridoxamine-P moiety from 315 to 325 nm, which increases the absorbance at the excitation wavelength for fluorescence. At pH 7, the absorption spectrum is already shifted and, consequently, the binding of PALA and carbamyl-P has little effect on the fluorescence. Therefore, the binding of these compounds at pH 8.0 must cause a structural change in the protein, which in turn causes protonation of a group in the modified active sites, altering the spectral properties.  相似文献   

10.
A comparison has been made of the values obtained by direct calorimetric measurements and van 't Hoff analysis, under similar conditions, for the enthalpy of binding of the bisubstrate analog N-phosphonacetyl-L-aspartate (PALA) to E. coli aspartate transcarbamylase and its catalytic subunit. In the case of the catalytic subunit, data were obtained at both saturating and non-saturating concentrations of L-Asp, and at two ionic strengths. Despite a 1000-fold difference in protein concentrations, and the obligatory omission of carbamyl phosphate in the calorimetric experiments, the values obtained by the two methods are shown to agree to within 15% when appropriate corrections are made. These results suggest that subunit dissociation is not a significant factor at the low protein concentrations used in the van 't Hoff analysis, and, conversely, that aggregation of the protein is negligible at the high protein concentrations used in the calorimetric experiments. They also imply that, at pH 8.3, the enthalpic difference between the two conformational states of the enzyme which exist in the presence and absence of substrates is less than 2.5 kcal/mol. In addition, the trends in the three sets of data for the catalytic subunit indicate that ionic bonds are involved in binding PALA to the active site, and that non-productive binding by L-Asp is negligible under these experimental conditions.  相似文献   

11.
Saturation curves of activity versus concentration were determined for aspartate transcarbamylase from Escherichia coli (EC 2.1.3.2) for the substrate L-aspartate at saturating carbamyl phosphate (4.8 mM) in buffered solution at pH values from 6.0 to 12.0. Hill coefficients were obtained from the sigmoidal curves. At pH values from 7.8 to 9.1, where substrate inhibition causes difficulties in the Hill approximation, our kinetic scheme includes substrate inhibition and residual activity in the abortive enzyme-substrate complex. The plot of Hill coefficient versus pH has pKalpha values of 7.4 and 9.8 at the half-maximum positions of the curve which has a plateau from pH 8.1 to 9.1. These pKalpha values may be associated with functional groups involved in the allosteric transition which activates the enzyme. A plot of [S]0.5 versus pH shows a pKalpha of 8.5, which may belong to a residue either at or near the aspartate binding site. At 50 mM aspartate concentration the pH-rate profile shows maxima at pH values of 8.8 and 10.0 (cf. Weitzman, P.D.J., and Wilson, I.B.(1966)J. Biol. Chem. 2418 5481-5488, who used 100 mM aspartate). However, when the pH-dependent substrate inhibition is included, the calculated Vmax--H curve is bell-shaped like that of the isolated catalytic subunit.  相似文献   

12.
In several experimental techniques D2O rather then H2O is often used as a solvent for proteins. Concerning the influence of the solvent on the stability of the proteins, contradicting results have been reported in literature. In this paper the influence of H2O-D2O solvent substitution on the stability of globular protein structure is determined in a systematic way. The differential scanning calorimetry technique is applied to allow for a thermodynamic analysis of two types of globular proteins: hen's egg lysozyme (LSZ) with relatively strong internal cohesion ("hard" globular protein) and bovine serum albumin (BSA), which is known for its conformational adaptability ("soft" globular protein). Both proteins tend to be more stable in D2O compared to H2O. We explain the increase of protein stability in D2O by the observation that D2O is a poorer solvent for nonpolar amino acids than H2O, implying that the hydrophobic effect is larger in D2O. In case of BSA the transitions between different isomeric forms, at low pH values the Nm and F forms, and at higher pH values Nm and B, were observed by the presence of a supplementary peak in the DSC thermogram. It appears that the pH-range for which the Nm form is the preferred one is wider in D2O than in H2O.  相似文献   

13.
The reaction of horse spleen ferritin (HoSF) with Fe2+ at pH 6.5 and 7.5 using O2, H2O2 and 1:1 a mixture of both showed that the iron deposition reaction using H2O2 is approximately 20- to 50-fold faster than the reaction with O2 alone. When H2O2 was added during the iron deposition reaction initiated with O2 as oxidant, Fe2+ was preferentially oxidized by H2O2, consistent with the above kinetic measurements. Both the O2 and H2O2 reactions were well defined from 15 to 40 degrees C from which activation parameters were determined. The iron deposition reaction was also studied using O2 as oxidant in the presence and absence of catalase using both stopped-flow and pumped-flow measurements. The presence of catalase decreased the rate of iron deposition by approximately 1.5-fold, and gave slightly smaller absorbance changes than in its absence. From the rate constants for the O2 (0.044 s(-1)) and H2O2 (0.67 s(-1)) iron-deposition reactions at pH 7.5, simulations of steady-state H2O2 concentrations were computed to be 0.45 microM. This low value and reported Fe2+/O2 values of 2.0-2.5 are consistent with H2O2 rapidly reacting by an alternate but unidentified pathway involving a system component such as the protein shell or the mineral core as previously postulated [Biochemistry 22 (1983) 876; Biochemistry 40 (2001) 10832].  相似文献   

14.
The ability of aspartate transcarbamylase from Escherichia coli to catalyze carbamylation of amino acids other than the natural substrate, L-aspartate, was examined. Cysteine, cysteate, cysteinesulfinate, and 3-nitroalanine showed kcat values at pH 7 of 0.16, 0.58, 5.2, and 62 s-1, respectively, while kcat with aspartate was 320 s-1. In a parallel study, competitive inhibition constants of 3-nitropropionate, 3-mercaptopropionate, 3-sulfopropionate, and 3-sulfinopropionate were found to be high, about 0.1 M, compared with that of succinate, 0.56 mM. Although cysteinesulfinate had low activity as a substrate, the pH dependences of kcat and kcat/Km in H2O and D2O observed with the compound closely paralleled those of aspartate. The results of these studies suggest that substrate specificity and reactivity are achieved in part by a strong, highly specific interaction of one or more active site residues with the beta-carboxylate of L-aspartate. Unlike the sigmoidal kinetics found with aspartate, saturation of native aspartate transcarbamylase by cysteine sulfinate showed a lack of cooperativity, even under conditions of activation of the reaction by ATP and inhibition by CTP. The cysteinesulfinate reaction was increased 9-fold by the bisubstrate analog N-phosphonacetyl-L-aspartate. These results were interpreted in terms of an inability of cysteinesulfinate to cause the allosteric conformational change promoted by aspartate.  相似文献   

15.
Ornithine transcarbamylase catalyzes the synthesis of citrulline from carbamyl phosphate and ornithine. This enzyme is involved in the biosynthesis of arginine in many organisms and participates in the urea cycle of mammals. The biosynthetic ornithine transcarbamylase has been purified from the filamentous fungus, Neurospora crassa. It was found to be a homotrimer with an apparent subunit molecular weight of 37,000 and a native molecular weight of about 110,000. Its catalytic activity has a pH optimum of 9.5 and Km's of about 5 and 2.5 mM for the substrates, ornithine and carbamyl phosphate, respectively, at pH 9.5. The Km's and pH optimum are much higher than those of previously characterized enzymes from bacteria, other fungi, and mammals. These unusual kinetic properties may be of significance with regard to the regulation of ornithine transcarbamylase in this organism, especially in the avoidance of a futile ornithine cycle. Polyclonal antibodies were raised against the purified enzyme. These antibodies and antibody raised against purified rat liver ornithine transcarbamylase were used to examine the structural similarities of the enzyme from a number of organisms. Cross-reactivity was observed only for mitochondrial ornithine transcarbamylases of related organisms.  相似文献   

16.
A study is presented on the pH dependence of proton translocation in the oxidative and reductive phases of the catalytic cycle of purified cytochrome c oxidase (COX) from beef heart reconstituted in phospholipid vesicles (COV). Protons were shown to be released from COV both in the oxidative and reductive phases. In the oxidation by O2 of the fully reduced oxidase, the H+/COX ratio for proton release from COV (R --> O transition) decreased from approximately 2.4 at pH 6.5 to approximately 1.8 at pH 8.5. In the direct reduction of the fully oxidized enzyme (O --> R transition), the H+/COX ratio for proton release from COV increased from approximately 0.3 at pH 6.5 to approximately 1.6 at pH 8.5. Anaerobic oxidation by ferricyanide of the fully reduced oxidase, reconstituted in COV or in the soluble case, resulted in H+ release which exhibited, in both cases, an H+/COX ratio of 1.7-1.9 in the pH range 6.5-8.5. This H+ release associated with ferricyanide oxidation of the oxidase, in the absence of oxygen, originates evidently from deprotonation of acidic groups in the enzyme cooperatively linked to the redox state of the metal centers (redox Bohr protons). The additional H+ release (O2 versus ferricyanide oxidation) approaching 1 H+/COX at pH < or = 6.5 is associated with the reduction of O2 by the reduced metal centers. At pH > or = 8.5, this additional proton release takes place in the reductive phase of the catalytic cycle of the oxidase. The H+/COX ratio for proton release from COV in the overall catalytic cycle, oxidation by O2 of the fully reduced oxidase directly followed by re-reduction (R --> O --> R transition), exhibited a bell-shaped pH dependence approaching 4 at pH 7.2. A mechanism for the involvement in the proton pump of the oxidase of H+/e- cooperative coupling at the metal centers (redox Bohr effects) and protonmotive steps of reduction of O2 to H2O is presented.  相似文献   

17.
Modification of Escherichia coli phosphofructokinase-2 (Pfk-2) with N-(1-pyrenil)maleimide results in an enzyme form that is inactive. However, the rate of modification is drastically reduced in the presence of the allosteric effector MgATP. The stoichiometry of the label incorporation was found to be 2.03 +/- 0.035 mol of the reagent/mol of subunit, in agreement with the number of titratable SH groups by 5,5'-dithiobis(2-nitrobenzoic acid) in the labeled protein. HPLC gel filtration experiments demonstrate that native Pfk-2 is a dimer in the absence of ligands, while in the presence of MgATP a dimer-tetramer transition is promoted. In contrast, the modified enzyme eluted as a monomer and the presence of MgATP was not able to induce aggregation. Although the modified monomers are inactive, the dissociation constants for the substrates and the allosteric effector MgATP, measured by following the fluorescence of the binding probe, are the same as for the native enzyme. Quenching of pyrene fluorescence emission of labeled phosphofructokinase-2 monomers by acrylamide gave downward curved Stern-Volmer plots, with very similar quenching efficiencies for the control and for the fructose-6-P and MgATP-enzyme complexes. These results show the presence of SH groups in the interface of Pfk-2 subunits, critical for subunit interactions, and that conformational changes occurring through the dimers are essential for catalytic activity.  相似文献   

18.
The role of conformational changes in the allosteric mechanism of aspartate transcarbamoylase from Escherichia coli was studied by reacting the isolated catalytic subunit with the bifunctional reagent tartryl diazide. Two derivatives differing moderately in substrate affinity were obtained depending on whether the reaction was conducted in the presence or absence of the substrate analogue succinate and carbamoyl phosphate. The modification was not accompanied by aggregation or dissociation. The modified catalytic subunits retained the ability to reassociate with unmodified regulatory subunits and produced hybrids similar in size to the native enzyme. These hybrids were appreciably sensitive to the allosteric effectors ATP and CTP but unlike native enzyme showed no cooperativity in substrate binding. The Michaelis constants of these hybrids for aspartate were intermediate between that of the isolated catalytic subunit and that of the relaxed state. Activation by ATP was caused by a reduction in Km to the value characteristic of the relaxed state whereas CTP inhibited by lowering the Vmax. The properties of the hybrids are strikingly similar to the modified enzyme obtained by Kerbiriou and Hervé from cells grown in the presence of 2-thiouracil. However, the crucial modifications are found in the regulatory subunits of the enzyme studied by these authors whereas they are located in the catalytic subunits of the hybrids reported here. Our results suggest that interactions between the catalytic and regulatory subunits have considerable effects on the state of the substrate binding sites in the native enzyme.  相似文献   

19.
ATP sulfurylase from Penicillium chrysogenum is a noncooperative homooligomer containing three free sulfhydryl groups per subunit. Under nondenaturing conditions, one SH group per subunit was modified by 5,5'-dithiobis-(2-nitrobenzoate), or N-ethylmaleimide. Modification had only a small effect on kcat, but markedly increased the [S]0.5 values for the substrates, MgATP and SO4(2-). MgATP and adenosine-5'-phosphosulfate protected against modification. The SH-modified enzyme displayed sigmoidal velocity curves for both substrates with Hill coefficients (nH) of 2. Fluorosulfonate (FSO3-) and other dead-end inhibitors competitive with SO4(2-) activated the SH-modified enzyme at low SO4(2-) concentration. In order to determine whether the sigmoidicity resulted from true cooperative binding (as opposed to a kinetically based mechanism), the shapes of the binding curves were established from the degree of protection provided by a ligand against phenylglyoxal-dependent irreversible inactivation under noncatalytic conditions. Under standard conditions (0.05 M Na-N-(2-hydroxyethyl)piperazine-N'-3-propanesulfonic acid buffer, pH 8, 30 degrees C, and 3mM phenylglyoxal) the native enzyme was inactivated with a k of 2.67 +/- 0.25 X 10-3 s-1, whereas k for the SH-modified enzyme was 5.44 +/- 0.27 X 10-3 s-1. The increased sensitivity of the modified enzyme resulted from increased reactivity of ligand-protectable groups. Both the native and the SH-modified enzyme displayed hyperbolic plots of delta k (i.e. protection) versus [MgATP], or [FSO3-], or [S2O3(2-]) in the absence of coligand (nH = 0.98 +/- 0.06). The plots of delta k versus [ligand] for the native enzyme were also hyperbolic in the presence of a fixed concentration of coligand. However, in the presence of a fixed [FSO3-] or [S2O3(2-]), the delta k versus [MgATP] plot for the SH-modified enzyme was sigmoidal, as was the plot of delta k versus [FSO3-] or [S2O3(2-]) in the presence of a fixed [MgATP]. The nH values were 1.92 +/- 0.09. The results indicate that substrates (or analogs) bind hyperbolically to unoccupied SH-modified subunits, but in a subunit-cooperative fashion to form a ternary complex.  相似文献   

20.
F Jordan  L Polgar  G Tous 《Biochemistry》1985,24(26):7711-7717
A technique was developed to exchange the backbone -N-H protons in D2O in the native subtilisins Carlsberg and BPN (Novo) that resulted in clearly resolved proton resonances in the aromatic region of the nuclear magnetic resonance spectrum. pH titration curves for four of the five histidine C2-H resonances in subtilisin Carlsberg and five of the six in subtilisin BPN between 7.5 and 8.8 ppm downfield from 4,4-dimethyl-4-silapentane-1-sulfonic acid sodium salt provided microscopic pKa's between 6.3 and 7.2 for both sources of the enzyme at ambient (approximately 22 degrees C) probe temperature. A resonance that titrated with a pKapp of 7.35 +/- 0.05 was observed in the 1H spectra only of the diisopropylphosphoryl derivatives of the subtilisins from both sources. The 31P NMR pH titration of the same preparations under identical conditions of solvent (D2O) and temperature gave a pKapp = 7.40 +/- 0.05 of the single titratable resonance. Both observations must pertain to His-64 at the active center. A resonance smaller than the others and titrating with a pKapp of 7.2 could also be observed in the native enzymes. This resonance was assigned to the catalytic center histidine since its pK corresponded to that derived from kinetic studies. No major perturbations in the chemical shifts or the pK's derived from the pH dependence of the observed resonances were apparent in the presence of saturating concentrations of the two putative transition-state analogues phenylboronic acid and bis [3,5-(trifluoromethyl)phenyl]boronic acid and in monoisopropylphosphorylsubtilisin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号