首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Isophthalic acid, 5-carboxy-, 5-hydroxy-, 5-methoxy-, 5-fluoro-, 5-bromo-, 5-cyano-, and 5-methylisophthalic acid were inhibitors competitive with L-glutamate for bovine liver glutamate dehydrogenase. The extent of inhibition by the derived compounds was not much greater than that obtained with the parent compound, isophthalic acid. A plot of pKi versus pH showed the presence of an ionizable group (pKa 7.4-7.8) at the enzyme active site which interacted with the substitutent at the 5 position of the substituted isophthalates.  相似文献   

2.
The yeast Saccharomyces cerevisiae takes up adenine, guanine, hypoxanthine, and cytosine via a common energy-dependent transport system. The apparent affinity of the transport system to these and other purines and pyrimidines is correlated with their capability to be protonated to the positively charged form. Further organic molecules are competitive inhibitors when they are cationic, e.g. guanidine and octylguanidine in contrast to urea, or hexadecyltrimethylammonium in contrast to dodecylsulfate and Triton X-100. The influence of the pH on the kinetic constants of hypoxanthine transport points to a stoichiometry of one proton being associated to the transport system together with one substrate molecule. The pKa values of two ionizable groups that are involved in substrate binding are revealed; one of which (pKa = 1.8) may be attributed to the substrate, the other (pKa = 5.1) to an amino acid residue in the recognition site of the transport system. Studies with group-specific inhibitors indicate that this amino acid residue contains a carboxyl group. The results are in accordance with the assumption that a carboxyl group of the transport system, a proton and a substrate molecule arrange to an uncharged ternary complex.  相似文献   

3.
W L Bigbee  F W Dahlquist 《Biochemistry》1977,16(17):3798-3803
The details of the pH dependence of the thermodynamic and magnetic interactions of the active-site region of thermolysin in which manganese has replaced the active-site zinc atom and the inhibitor N-trifluoroacetyl-D-phenylalanine have been examined. These show a number of ionizable groups in the active-site region. A cooperative displacement of manganese at the catalytic site is observed as pH is lowered. This appears to be the result of the protonation of histidine-142 and -146 which act as metal ligands. The metal is 50% displaced at pH 6.0. At higher pH values, the environment of the bound manganese changes as a result of the ionization of at least two groups of approximate pKa = 8.5 and 9.5. These values are assigned to tyrosine-157 and to the water molecule which acts as a metal ligand at the active site. The binding behavior of the inhibitor strongly suggests that two molecules of inhibitor bind to the enzyme. The weaker site is competitive with the synthetic substrate FAGLA (furylacryloylglycyl-leucinamide), while the strong site has no effect on FAGLA hydrolysis. This second site is in the vicinity of the active site with a distance of 8 A or less between the trifluoromethyl group and manganese bound at the active site.  相似文献   

4.
Glutamine-synthetase (GS) from Beta vulgaris seedlings, purified 150-fold, was characterized with regard to its physiological substrate NH3. The data were compared to the unphysiological substrate NH2−OH frequently used in the assay (both synthetase and transferase reaction). The pH-optimum was found at pH 7.5 for the synthetase and at pH 6.3 for the transferase reaction. Through plots of pKm vs pH, the pKe values for dissociable groups in the reaction center were found to be in the range from pH 7–8. Mg2+-ion serves as an allosteric effector with a Hill coefficient of 4.2. The results are discussed in relation to the control of nitrogen metabolism in Beta .  相似文献   

5.
When the pKm of alpha-mannosidase was determined at different pH values, the results indicated that ionizable groups with pK values of approx. 3.8 and 5.7 could be essential. Modification with carbodiimide or Woodward's Reagent K abolished the enzyme activity. The substrate analogue, alpha-methyl-D-mannoside, protected the enzyme against inactivation. Incorporation of a 14C-labeled nucleophile reagent in the presence or absence of the analogue suggested that 2--4 carboxyl groups were protected. Exchange studies indicated that the essential Zn2+ could be bound to such groups. There was no indication that hydroxyl groups, sulphydryl groups, guanidino groups or amino groups take part in the catalytic activity.  相似文献   

6.
Chemical modification studies were performed to elucidate the role of Cys-residues in the catalysis/binding of restriction endonuclease Cfr9I. Incubation of restriction endonuclease Cfr9I with N-ethylmaleimide (NEM), iodoacetate, 5,5'-dithiobis (2-nitrobenzoic acid) at pH 7.5 led to a complete loss of the catalytic activity. However, no enzyme inactivation was detectable after modification of the enzyme with iodoacetamide and methyl methanethiosulfonate. Complete protection of the enzyme against inactivation by NEM was observed in the presence of substrate implying that Cys-residues may be located at or in the vicinity of the active site of enzyme. Direct substrate-binding studies of native and modified restriction endonuclease Cfr9I using a gel-mobility shift assay indicated that the modification of the enzyme by NEM was hindered by substrate binding. A single Cys-residue was modified during the titration of the enzyme with DTNB with concomitant loss of the catalytic activity. The pH-dependence of inactivation of Cfr9I by NEM revealed the modification of the residue with the pKa value of 8.9 +/- 0.2. The dependence of the reaction rate of substrate hydrolysis by Cfr9I versus pH revealed two essential residues with pKa values of 6.3 +/- 0.15 and 8.7 +/- 0.15, respectively. The evidence presented suggests that the restriction endonuclease Cfr9I contains a reactive sulfhydryl residue which is non-essential for catalysis, but is located at or near the substrate binding site.  相似文献   

7.
The nitro analogs of aspartate and argininosuccinate were synthesized and tested as substrates and inhibitors of argininosuccinate synthetase and argininosuccinate lyase, respectively. The Vmax for 3-nitro-2-aminopropionic acid was found to be 60% of the maximal rate of aspartate utilization in the reaction catalyzed by argininosuccinate synthetase. Only the nitronate form of this substrate, in which the C-3 hydrogen is ionized, was substrate active, indicating a requirement for a negatively charged group at the beta carbon. The V/K of the nitro analog of aspartate was 85% of the value of aspartate after correcting for the percentage of the active nitronate species. The nitro analog of argininosuccinate, N3-(L-1-carboxy-2-nitroethyl)-L-arginine, was a strong competitive inhibitor of argininosuccinate lyase but was not a substrate. The pH dependence of the observed pKi was consistent with the ionized carbon acid (pK = 8.2) in the nitronate configuration as the inhibitory material. The pH-independent pKi of 2.7 microM is 20 times smaller than the Km of argininosuccinate at pH 7.5. These results suggest that the tighter binding of the nitro analog relative to the substrate is due to the similarity in structure to a carbanionic intermediate in the reaction pathway.  相似文献   

8.
The active site amino acid residues of lignin peroxidase are homologous to those of other peroxidases; however, in contrast to other peroxidases, no pH dependence is observed for the reaction of ferric lignin peroxidase with H2O2 to form compound I (Andrawis, A., Johnson, K.A., and Tien, M. (1988) J. Biol. Chem. 263, 1195-1198). Chloride binding is used in the present study to investigate this reaction further. Chloride binds to lignin peroxidase at the same site as cyanide and hydrogen peroxide. This is indicated by the following. 1) Chloride competes with cyanide in binding to lignin peroxidase. 2) Chloride is a competitive inhibitor of lignin peroxidase with respect to H2O2. The inhibition constant (Ki) is equal to the dissociation constant (Kd) of chloride at all pH values studied. Chloride binding is pH dependent: chloride binds only to the protonated form of lignin peroxidase. Transient-state kinetic studies demonstrate that chloride inhibits lignin peroxidase compound I formation in a pH-dependent manner with maximum inhibition at low pH. An apparent pKa was calculated at each chloride concentration; the pKa increased as the chloride concentration increased. Extrapolation to zero chloride concentration allowed us to estimate the intrinsic pKa for the ionization in the lignin peroxidase active site. The results reported here provide evidence that an acidic ionizable group (pKa approximately 1) at the active site controls both lignin peroxidase compound I formation and chloride binding. We propose that the mechanism for lignin peroxidase compound I formation is similar to that of other peroxidases in that it requires the deprotonated form of an ionizable group near the active site.  相似文献   

9.
The inhibition of purified bovine adrenal tyrosine hydroxylase by several product and substrate analogues has been studied to probe the kinetic mechanism. Norepinephrine, dopamine, and methylcatechol are competitive inhibitors versus tetrahydropterins and noncompetitive inhibitors versus tyrosine. 3-Iodotyrosine is an uncompetitive inhibitor versus tetrahydropterins and a competitive inhibitor versus tyrosine. The Ki value for 3-iodotyrosine depends on the tetrahydropterin used. These results are consistent with tetrahydropterin binding first to the free enzyme followed by binding of tyrosine. 5-Deaza-6-methyltetrahydropterin is a noncompetitive inhibitor versus tetrahydropterins and tyrosine. The effect of varying the concentration of tyrosine on the Ki value for 5-deaza-6-methyltetrahydropterin is consistent with the binding of this inhibitor to both the free enzyme and to an enzyme-dihydroxyphenylalanine complex. Dihydroxyphenylalanine also is a noncompetitive inhibitor versus tetrahydropterins and tyrosine; the effect of changing the fixed substrate is consistent with the binding of this inhibitor to both the free enzyme and to the enzyme-tetrahydropterin complex. The effect of pH on the Ki values was determined in order to measure the pKa values of amino acid residues involved in substrate binding. Tight binding of catechols requires that a group with a pKa value of 7.6 be deprotonated. Binding of 3-iodotyrosine involves two groups with pKa values of 7.5 and about 5.5, one of which must be protonated for binding. Binding of 5-deaza-6-methyltetrahydropterin requires that a group on the free enzyme with a pKa value of 6.1 be protonated. The Ki value for dihydroxyphenylalanine is relatively insensitive to pH, but the inhibition pattern changes from noncompetitive to competitive above pH 7.5, consistent with the measured pKa values for binding to the free enzyme and to the enzyme-tetrahydropterin complex.  相似文献   

10.
Cathepsin B has been shown to catalyze the transfer of the N alpha-benzyloxycarbonyl-L-lysyl residue from the corresponding p-nitrophenyl ester substrate to water and dipeptide nucleophiles. These reactions occurred through the formation of an acyl-enzyme intermediate. The pH dependency of the acylation and deacylation steps were determined from the increases in the maximum rate of appearance of p-nitrophenol on addition of glycylglycine or L-leucylglycine to the reaction. The second order acylation rate constant, kcat/Km was found to depend on the state of ionization of three groups in the enzyme having pKa values of 4.2, 5.5, and 8.6. Protonation of the group with pKa = 5.5 decreased but did not abolish enzymatic activity, resulting in the appearance of a second, active protonic form of the enzyme between pH 4.2 and pH 5.5. The first order rate constant for the hydrolysis of the acyl-enzyme intermediate was independent of pH between 4.0 and 7.5. In contrast, acyl group transfer from cathepsin B to glycylglycine and L-leucylglycine depended on a group with a pKa of about 4.5. These results are discussed in terms of possible structural and functional homologies between the active sites of cathepsin B and papain.  相似文献   

11.
A kinetic study has been carried out over the pH range of 2.63-9.37 for the reaction of horseradish peroxidase with hydrogen peroxide to form compound I of th;e enzyme. Analysis of the results, indicates that there are two kinetic influencing, ionizable groups on the enzyme with pKa values of 3.2 and 3.9. Protonation of these groups results in a decrease in the rate of reaction of the enzyme with H2O2. A previous study of the kinetics of cyanide binding to horseradish peroxidase (Ellis, W.D. & Dunford, H.B.: Biochemistry 7, 2054-2062 (1968)) has been extended to down to pH 2.55, and analysis of these results also indicates the presence of two kinetically important ionizable groups on the enzyme with pKa values of 2.9 and 3.9.  相似文献   

12.
Yunhui Peng  Emil Alexov 《Proteins》2017,85(2):282-295
Protein–nucleic acid interactions play a crucial role in many biological processes. This work investigates the changes of pKa values and protonation states of ionizable groups (including nucleic acid bases) that may occur at protein–nucleic acid binding. Taking advantage of the recently developed pKa calculation tool DelphiPka, we utilize the large protein–nucleic acid interaction database (NPIDB database) to model pKa shifts caused by binding. It has been found that the protein's interfacial basic residues experience favorable electrostatic interactions while the protein acidic residues undergo proton uptake to reduce the energy cost upon the binding. This is in contrast with observations made for protein–protein complexes. In terms of DNA/RNA, both base groups and phosphate groups of nucleotides are found to participate in binding. Some DNA/RNA bases undergo pKa shifts at complex formation, with the binding process tending to suppress charged states of nucleic acid bases. In addition, a weak correlation is found between the pH‐optimum of protein–DNA/RNA binding free energy and the pH‐optimum of protein folding free energy. Overall, the pH‐dependence of protein–nucleic acid binding is not predicted to be as significant as that of protein–protein association. Proteins 2017; 85:282–295. © 2016 Wiley Periodicals, Inc.  相似文献   

13.
The observation that 6 M-urea denatures horse ferricytochrome c in the pH range 4-6, but not horse ferrocytochrome c, has been exploited to determine the denaturation-induced proton uptake of ferricytochrome c. This is related to the pKa values of ionizable groups buried within the native protein. The data indicate that one of the haem propionic acid substituents of ferricytochrome c has a pKa of less than 4.5, whereas the other has a pKa of greater than 9.  相似文献   

14.
The concentration dependence of the rate of hydrolysis of L-asparagine by Escherichia coli L-asparaginase (L-asparagine amidohydrolase, EC 3.5.1.1) has been measured over the range pH 4.5 to pH 9.1 by a direct spectrophotometric assay at 220 nm and by a coupled assay utilizing glutamate dehydrogenase to detect the ammonia produced. The velocity of the hydrolysis reaction at saturating levels of substrate is independent of pH over this interval. The plot of V/km over the same interval is bell-shaped, being dependent on pKa values of 6.58 and 8.69. The higher pKa is attributed to the amino group of asparagine. The lower pKa is associated with the enzyme active site and is probably due to an imidazole group.  相似文献   

15.
The pH dependence of the two-dimensional 1H nuclear magnetic resonance spectra of hen and turkey egg-white lysozymes has been recorded over the pH range 1-7. By monitoring the chemical shifts of the resonances of the various protons of ionizable residues, individual pKa values for the acidic residues have been determined for both proteins. The pKa values are displaced, with the exception of those of the residues in the active site cleft, by an average of 1 unit to low pH compared to model compounds.  相似文献   

16.
1H NMR spectroscopy of the isotropically shifted signals in cobalt carboxypeptidase, CoCPD, permits a direct and selective detection of protons belonging to the residues liganded to the metal. The chemical shift of these protons in the free enzyme and enzyme-inhibitor complexes with changing pH monitors the state of ionization of the ligands directly and of other residues in the active center indirectly. The 1H NMR spectrum of CoCPD at pH 6 shows three well-resolved isotropically shifted signals in the downfield region at 62 (a), 52 (c), and 45 (d) ppm which have been assigned to the NH proton of His-69 and to the C-4 H's of His-69 and His-196, respectively. Titration of signal a with pH is characterized by a pKa of 8.8 which is identical to that seen in prior electronic absorption and kinetic studies. The fact that the signal reflecting the NH of His-69 is still observed at pH 10 and no major shifts occur for the signals reflecting the C-4 H's indicates the alkaline pKa in carboxypeptidase A catalysis, pKEH, cannot be ascribed to ionization of the histidyl NH of either His-69 or His-196. Binding of L-Phe shifts this pKa to 7.7 while not greatly perturbing the downfield 1H NMR signals that reflect the ligation shell of the cobalt coordination sphere. These results indicate the pKa of 8.8 in CoCPD and the pKa of 7.7 in the CoCPD.L-Phe adduct reflect ionization of the same group.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The pH dependence of the kinetics of the binding of cyanide ion to methemoglobins A and S and to guinea pig and pigeon methemoglobins appears to be not directly correlated with the net charges on the proteins. The kinetics can, however, be adequately explained in terms of three sets of heme-linked ionizable groups with pK1 ranging between 4.9 and 5.3, pK2 between 6.2 and 7.9, and pK3 between 8.0 and 8.5 at 20 degrees C. pK1 is assigned to carboxylic acid groups, pK2 to histidines and terminal amino groups, and pK3 to the acid-alkaline methemoglobin transition. Kinetic second order rate constants have also been determined for the binding of cyanide ion by the four sets of methemoglobin species present in solution. The pKi values and the rate constants of methemoglobin S are strikingly different from those of methemoglobin A. This result is explained in terms of different electrostatic contributions to the free energy of heme linkage arising from differences in the environments of ionizable groups at the surfaces of the two molecules.  相似文献   

18.
Qian J  West AH  Cook PF 《Biochemistry》2006,45(39):12136-12143
Homocitrate synthase (acetyl-coenzyme A:2-ketoglutarate C-transferase; E.C. 2.3.3.14) catalyzes the condensation of AcCoA and alpha-ketoglutarate to give homocitrate and CoA. The enzyme was found to be a Zn-containing metalloenzyme using inductively coupled plasma mass spectrometry. Dead-end analogues of alpha-ketoglutarate were used to obtain information on the topography of the alpha-ketoglutarate binding site. The alpha-carboxylate and alpha-oxo groups of alpha-ketoglutarate are required for optimum binding to coordinate to the active site Zn. Optimum positioning of the alpha-carboxylate, alpha-oxo, and gamma-carboxylate of alpha-ketoglutarate is likely mimicked by the location in space of the 2-carboxylate, pyridine nitrogen, and 4 carboxylate of pyridine 2,4-dicarboxylate. The pH dependence of the kinetic parameters was determined to obtain information on the chemical mechanism of homocitrate synthase. The V profile is bell shaped with slopes of 1 and -1, giving pKa values of 6.7 and 8.0, while V/K(AcCoA) exhibits a slope of 2 on the acidic side with an average pKa value of 6.6 and a slope of -2 on basic side of the profile with an average pKa value of 8.2. The V/K(alpha-Kg) pH-rate profile exhibits a single pKa of 6.9 on the acidic side and two on the basic side with an average value of 7.8. The pH dependence of the Ki for glyoxylate, a competitive inhibitor vs alpha-ketoglutarate, gives a pKa of 7.1 for a group, required to be protonated for optimum binding. Data suggest a chemical mechanism for the enzyme in which alpha-ketoglutarate first binds to the active site Zn via its alpha-carboxylate and alpha-oxo groups, followed by acetyl-CoA. A general base then accepts a proton from the methyl of acetyl-CoA, and a general acid protonates the carbonyl of alpha-ketoglutarate in the formation of homocitryl-CoA. The general acid then acts as a base in deprotonating Zn-OH2 in the hydrolysis of homocitryl-CoA to give homocitrate and CoA. A solvent deuterium kinetic isotope effect of 1 is measured for homocitrate synthase, while a small pH-independent primary kinetic deuterium isotope effect (approximately 1.3) is observed using deuterioacetyl-CoA. Data suggest rate-limiting condensation to form the alkoxide of homocitryl-CoA, followed by hydrolysis to give products.  相似文献   

19.
Partial amino acid sequences, the essential ionizable groups directly involved in catalytic reaction, and the subsite structure of beta-D-glucosidase purified from a Streptomyces sp. were investigated in order to analyze the reaction mechanism. On the basis of the partial amino acid sequences, the enzyme seemed to belong to the family 1 of beta-glucosidase in the classification of glycosyl hydrolases by Henrissat (1991). Dependence of the V and Km values on pH, when the substrate concentration was sufficiently lower than Km, gave the values of 4.1 and 7.2 for the ionization constants, pKe1 and pKe2 of essential ionizable groups 1 and 2 of the free enzyme, respectively. When the dielectric constant of the reaction mixture was decreased in the presence of 10% methanol, the pKe1 and pKe2, values shifted to higher, to +0.60 and +0.35 pH unit, respectively. The findings supported the notion that the essential ionizable groups of the enzyme were a carboxylate group (-COO-, the group 1) and a carboxyl group (-COOH, the group 2). The subsite affinities Ai's in the active site were evaluated on the basis of the rate parameters of laminarioligosaccharides. Subsites 1 and 2 having positive Ai values (A1 was 1.10 kcal/mol and A2 was 4.98 kcal/mol) were considered to probably facilitate the binding of the substrate to the active site. However, the subsites 3 and 4 showed negative Ai values (A3 was -0.21 kcal/mol and A4 was -2.8 kcal/mol).  相似文献   

20.
Mechanistic studies on thrombin catalysis   总被引:1,自引:0,他引:1  
S R Stone  A Betz  J Hofsteenge 《Biochemistry》1991,30(41):9841-9848
The kinetic mechanism of the cleavage of four p-nitroanilide (pNA) substrates by human alpha-thrombin has been investigated by using a number of steady-state kinetic techniques. Solvent isotope and viscosity effects were used to determine the stickiness of the substrates at the pH optimum of the reaction; a sticky substrate is defined as one that undergoes catalysis faster than it dissociates from the Michaelis complex. Whereas benzoyl-Arg-pNA could be classified as a nonsticky substrate, D-Phe-pipecolyl-Arg-pNA was very sticky. The other two substrates (tosyl-Gly-Pro-Arg-pNA and acetyl-D-Phe-pipecolyl-Arg-pNA) were slightly sticky. The pH profiles of kcat/Km were bell-shaped for all substrates. The pKa values determined from the pH dependence of kcat/Km for benzoyl-Arg-pNA were about 7.5 and 9.1. Similar pKa values were determined from the pH profiles of kcat/Km for tosyl-Gly-Pro-Arg-pNA and acetyl-D-Phe-pipecolyl-Arg-pNA and for the binding of the competitive inhibitor N alpha-dansyl-L-arginine-4-methylpiperidine amide. The groups responsible for the observed pKa values were proposed to be His57 and the alpha-amino group of Ile16. The temperature dependence of the pKa values was consistent with this assignment. The pKa values of 6.7 and 8.6 observed in the pH profile of kcat/Km for D-Phe-pipecolyl-Arg-pNA were displaced to lower values than those observed for the other substrates. The displacement of the acidic pKa value could be attributed to the stickiness of this substrate.(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