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.
J Pohl  B M Dunn 《Biochemistry》1988,27(13):4827-4834
The possibility that pig pepsin has a cation binding specificity in its secondary binding subsites has been examined by the pepsin-catalyzed hydrolysis of a series of synthetic octa- to undecapeptide substrates. These chromophoric substrates are cleaved by pepsin in the phenylalanyl-p-nitrophenylalanyl (Phe-Nph) bond. Lys and Arg residues were placed into seven different positions in the substrates, and their effect on kcat and Km was examined between pH 2.8 and pH 5.8 (I = 0.1 M, 37 degrees C). Kinetic evidence indicates the existence in the enzyme binding subsites S4, S3, S2, S3', S4', and S5' of a group(s) which become(s) negatively charged at higher pH. For most substrates, the magnitude as well as the pH dependence of kcat was unaffected by the presence of Lys or Arg in these peptides. In contrast, changes up to 5 orders of magnitude were observed for Km, depending on the number of basic residues and on their positions in the sequence. Km for a group of substrates at pH greater than 5.5 was lower than 50 nM. Values for kcat/Km for some substrates exceed the level of 10(8) M-1 s-1. Therefore, the free energy derived from ionic interactions in secondary binding sites influences mostly the binding step on the reaction pathway. This result is in contrast to the previous observations that the length and the hydrophobic character of the substrate residues in some positions influence kcat with little effect on Km toward shorter substrates of pepsin [Fruton, J. (1976) Adv. Enzymol. Relat. Areas Mol. Biol. 44, 1-36].  相似文献   

2.
Nitrogenase from the facultative anaerobe Bacillus polymxa was separated into its component proteins, which were recombined in the ratio that produced optimal specific activity (125 to 175 nmol of C2H2 reduced/min per mg of total protein). The apparent Michaelis constants (Km)for the magnesium adenosine triphosphate complex, reducible substrates azide, acetylene, and N2 and the nonphysiological electron donor hydrosulfite (S2O42-) were determined to be 0.7, 0.7, 0.2, 0.06, and 0.03 MM, respectively. These apparent Km values are in reasonable agreement with those reported for the nitrogenases of Azotobacter vinelandii and Klebsiella pneumoniae. Either a total lack of cooperativity between binding sites or a single binding site for reducible substrates is indicated by analysis of Hill plots. Hill plot slopes of approximately 1.7 suggest that multiple binding sites exist for both ATP and S2O42-.  相似文献   

3.
The kinetic and regulatory properties of cGMP-activated phosphodiesterase (PDE) from human brain were studied. In double reciprocal plots the enzyme activity is characterized by a linear dependence of cAMP and a nonlinear one for cGMP. Micromolar concentrations of cGMP accelerate cAMP hydrolysis (7-14-fold) with Ka for cGMP of 0.36 microM. Stimulation of cAMP hydrolysis is accompanied by a decrease of Km with no changes in Vmax. With a rise in the cGMP concentration above 5 microM PDE activation is changed by its inhibition. Both substrates act as competitive inhibitors towards each other. The Ki value for both cGMP and cAMP is 30 microM. After the increase in the cAMP (Bt)2 concentration the activation of 5 microM cAMP hydrolysis is accompanied by the enzyme inhibition. Both analogs competitively inhibit cGMP hydrolysis with Ki of 10 and 1500 microM for cGMP(Bt)2 and cAMP(Bt)2, respectively. The data obtained point to the existence of two binding sites for cyclic nucleotides, namely, a regulatory site which is highly specific for cGMP and a catalytic site responsible for the hydrolysis of the both substrates which displays no apparent specificity either for cAMP or for cGMP. The different affinity of natural and synthetic cyclic nucleotides for these sites is determined, to a large extent, by the amino groups in the 2nd and 6th positions of the purine ring.  相似文献   

4.
Cathepsin C, a lysosomal dipeptidyl aminopeptidase, is competitively and reversibly inhibited by guanidinium ions with a Ki approximately 1.5 mM. Loss of activity is not the result of conformational change, subunit dissociation or altered mobility of the enzyme, but rather reflects a specific binding of guanidinium ions to the active site. The finding that cathepsin C is not inhibited by substrate has allowed the kinetic parameters in the presence of guanidinium ion to be determined. Guanidinium significantly decreases the Km of substrate hydrolysis, without changing Vmax. In a novel application of the transferase reaction, the Km of the nucleophile substrate has been determined (11 mM) and found not to be affected by guanidinium, indicating its inhibition of substrate binding to the S, but not the S', site. Inhibition is suggested to be the result of shielding a negative charge on the enzyme important for interaction with the substrate.  相似文献   

5.
1. Several peptides containing either of the sequences -Phe(NO2)-Trp- and -Phe(NO2)-Phe- and an uncharged hydrophilic group were synthesized, and the steady-state kinetics of their hydrolysis by pig pepsin (EC 3.4.23.1) and chicken liver cathepsin D (EC 3.4.23.5) were determined. Despite the presence of a hydrophilic group to increase substrate solubility, it was not possible to achieve the condition [S]0 much greater than Km, and, in some cases, only values of kcat./Km could be determined by measuring the first-order rate constant when [S]0 much less than Km. 2. Occupancy of the P2 and P3 sites considerably enhanced the specificity constant, and alanine was more effective than glycine at site P2. 3. The specificity constants for the hydrolysis by pepsin of those substrates in the present series that contain an amino acid residue at site P3 are considerably lower than for comparable substrates containing a cationic group. This difference does not apply to cathepsin D. 4. Hydrolyses with cathepsin D commonly exhibited a lag phase, and a possible explanation for this is given.  相似文献   

6.
The effects of tyramine, serotonin and benzalkonium on the esterase and aryl acylamidase activities of wild-type human butyrylcholinesterase and its peripheral anionic site mutant, D70G, were investigated. The kinetic study was carried out under steady-state conditions with neutral and positively charged aryl acylamides [o-nitrophenylacetanilide, o-nitrotrifluorophenylacetanilide and m-(acetamido) N,N,N-trimethylanilinium] and homologous esters (o-nitrophenyl acetate and acetylthiocholine). Tyramine was an activator of hydrolysis for neutral substrates and an inhibitor of hydrolysis for positively charged substrates. The affinity of D70G for tyramine was lower than that of the wild-type enzyme. Tyramine activation of hydrolysis for neutral substrates by D70G was linear. Tyramine was found to be a pure competitive inhibitor of hydrolysis for positively charged substrates with both wild-type butyrylcholinesterase and D70G. Serotonin inhibited both esterase and aryl acylamidase activities for both positively charged and neutral substrates. Inhibition of wild-type butyrylcholinesterase was hyperbolic (i.e. partial) with neutral substrates and linear with positively charged substrates. Inhibition of D70G was linear with all substrates. A comparison of the effects of tyramine and serotonin on D70G versus the wild-type enzyme indicated that: (a) the peripheral anionic site is involved in the nonlinear activation and inhibition of the wild-type enzyme; and (b) in the presence of charged substrates, the ligand does not bind to the peripheral anionic site, so that ligand effects are linear, reflecting their sole interaction with the active site binding locus. Benzalkonium acted as an activator at low concentrations with neutral substrates. High concentrations of benzalkonium caused parabolic inhibition of the activity with neutral substrates for both wild-type butyrylcholinesterase and D70G, suggesting multiple binding sites. Benzalkonium caused linear, noncompetitive inhibition of the positively charged aryl acetanilide m-(acetamido) N,N,N-trimethylanilinium for D70G, and an unusual mixed-type inhibition/activation (alpha > beta > 1) for wild-type butyrylcholinesterase with this substrate. No fundamental difference was observed between the effects of ligands on the butyrylcholinesterase-catalysed hydrolysis of esters and amides. Thus, butyrylcholinesterase uses the same machinery, i.e. the catalytic triad S198/H448/E325, for the hydrolysis of both types of substrate. The differences in response to ligand binding depend on whether the substrates are neutral or positively charged, i.e. the differences depend on the function of the peripheral site in wild-type butyrylcholinesterase, or the absence of its function in the D70G mutant. The complex inhibition/activation effects of effectors, depending on the integrity of the peripheral anionic site, reflect the allosteric 'cross-talk' between the peripheral anionic site and the catalytic centre.  相似文献   

7.
The rate constants and Km for the hydrolysis of the optically active nonglycosidic analogues of the CpA and C greater than p catalysed by RNase A and RNase BS-I were measured. The rate of hydrolysis of the model substrates in 10(5) and 10(3) slower that for the appropriate dinucleoside phosphate and nucleoside cyclophosphate. However, substitution of the relatively rigid ribofuranose ring with flexible alifatic chains is accompanied by little variation in binding constants. The analyses based on the single substrate system indicate that the observed difference in rate constants must be accounted for by a difference between the binding of the substrates in the transition state to the RNase active site. Consequently, the "rigidity" of the ribose rings in RNA leads to large decreases in the free energy of activation for the reactions catalysed by RNases.  相似文献   

8.
The rates of hydrolysis of the ester, amide and anilide substrates of p-guanidino-L-phenylalanine (GPA) by Streptomyces griseus trypsin (S. griseus trypsin) were compared with those of arginine (Arg) substrates. The specificity constant (kcat/km) for the hydrolysis of GPA substrates by the enzyme was 2-3-times lower than that for arginine substrates. The kcat and Km values for the hydrolysis of N alpha-benzoyl-p-guanidino-L-phenylalanine ethyl ester (Bz-GPA-OEt) by S. griseus trypsin are in the same order of magnitude as those of N alpha-benzoyl-L-arginine ethyl ester (Bz-Arg-OEt), although both values for the former when hydrolyzed by bovine trypsin are higher by one order of magnitude than those for the latter. The specificity constant for the hydrolysis of Bz-GPA-OEt by S. griseus trypsin is much higher than that for N alpha-benzoyl-p-guanidino-L-phenylglycine ethyl ester (Bz-GPG-OEt). As with the kinetic behavior of bovine trypsin, low values in Km and kcat were observed for the hydrolysis of amide and anilide substrates of GPA by S. griseus trypsin compared with those of arginine substrates. The rates of hydrolysis of GPA and arginine substrates by S. griseus trypsin are about 2- to 62-times higher than those obtained by bovine trypsin. Substrate activation was observed with S. griseus trypsin in the hydrolysis of Bz-GPA-OEt as well as Bz-Arg-OEt, whereas substrate inhibition was observed in three kinds of N alpha-protected anilide substrates of GPA and arginine. In contrast, no activation by the amide substrate of GPA could be detected with this enzyme.  相似文献   

9.
A serine proteinase (ycaB) from the yeast Candida albicans A.T.C.C. 10261 was purified to near homogeneity. The enzyme was almost indistinguishable from yeast proteinase B (EC 3.4.21.48), and an Mr of 30,000 for the proteinase was determined by SDS/polyacrylamide-gel electrophoresis. The initial site of hydrolysis of the oxidized B-chain of insulin, by the purified proteinase, was the Leu-Tyr peptide bond. The preferential degradation at this site, analysed further with N-blocked amino acid ester and amide substrates, demonstrated that the specificity of the proteinase is determined by an extended substrate-binding site, consisting of at least three subsites (S1, S2 and S'1). The best p-nitrophenyl ester substrates were benzyloxycarbonyl-Tyr p-nitrophenyl ester (kcat./Km 3,536,000 M-1 X S-1), benzyloxycarbonyl-Leu p-nitrophenyl ester (kcat./Km 2,250,000 M-1 X S-1) and benzyloxycarbonyl-Phe p-nitrophenyl ester (kcat./Km 1,000,000 M-1 X S-1) consistent with a preference for aliphatic or aromatic amino acids at subsite S1. The specificity for benzyloxycarbonyl-Tyr p-nitrophenyl ester probably reflects the binding of the p-nitrophenyl group in subsite S'1. The presence of S2 was demonstrated by comparison of the proteolytic coefficients (kcat./Km) for benzyloxycarbonyl-Ala p-nitrophenyl ester (825,000 M-1 X S-1) and t-butyloxycarbonyl-Ala p-nitrophenyl ester (333,000 M-1 X S-1). Cell-free extracts contain a heat-stable inhibitor of the proteinase.  相似文献   

10.
To assess the relative importance of binding to enzyme-substrate complex (E.S) and to acetylenzyme (EA), noncompetitive inhibition has been studied in hydrolysis by acetylcholinesterase (AcChE) of cationic and uncharged substrates - acetylcholine (AcCh), 3,3-dimethylbutyl acetate, n-butyl acetate, 2-(methylammonio)ethyl acetate, 2- (N,N-diethyl-N-n-butylammonio)ethyl acetate (DEBAAc) and 2-(methylsulfonyl)ethyl acetate. For the N-trimethyl quaternary ions related to AcCh, tetramethylammonium ion, choline and choline ethyl ether, noncompetitive inhibition (Ki(nonc) is more favorable with the slower substrates than with AcCh, i.e., when E.S greater than EA, and is attributed to formation of enzyme-substrate-inhibitor complexes, E.S.I'. Noncompetitive inhibition by tetraethyl-, tert-butyl- and isopropylammonium ions, and acetamidocholine and its lower dimethyl analogue, is also attributed to E.S.I' complexes. Peripheral binding of these inhibitors decreases acylation more than deacylation. Some tertiary dimethylamonio ions have more favorable Ki(nonc) values with AcCh, decreasing deacylation more than acylation. The substrate DEBAAc is a more effective noncompetitive than competitive inhibitor in hydrolysis of AcCh, indicating that it binds more strongly in a peripheral site than in the active site of the free enzyme. In its hydrolysis by AcChE, it acts as its own noncompetitive inhibitor, by this non-productive binding. Formation of E.S.I' complexes is a general characteristic of hydrolysis by AcChE and decrease in rates at high concentrations of AcCh and related substrates is attributed to peripheral regulatory site binding, formation of E.S.S' complexes, rather than to binding to the acetylenzyme.  相似文献   

11.
The initial rates of hydrolysis of Bz-Gly-Lys and Bz-Gly-Phe by carboxypeptidase B (CPB) are increased in the presence of the modifiers beta-phenylpropionic acid, cyclohexanol, Bz-Gly, and Bz-Gly-Gly. The hydrolysis of the tripeptide Bz-Gly-Gly-Phe is also activated by Bz-Gly and Bz-Gly-Gly, but none of these modifiers activate the hydrolysis of Bz-Gly-Gly-Lys, Z-Leu-Ala-Phe, or Bz-Gly-phenyllactic acid by CPB. All modifiers except cyclohexanol display inhibitory modes of binding when present in high concentration. Examination of Lineweaver-Burk plots in the presence of fixed concentrations of Bz-Gly has shown that activation of the hydrolysis of neutral and basic peptides by CPB, as reflected in the values of the extrapolated parameters, Km(app) and kcat, occurs by different mechanisms. For Bz-Gly-Gly-Phe, activation occurs because the enzyme-modifier complex has a higher affinity than the free enzyme for the substrate, whereas activation of the hydrolysis of Bz-Gly-Lys derives from an increase in the rate of breakdown of the enzyme-substrate complex to give products. Cyclohexanol differs from Bz-Gly and Bz-Gly-Gly in that it displays no inhibitory mode of binding with any of the substrates examined, activates only the hydrolysis of dipeptides by CPB, and has a greater effect on the hydrolysis of the basic dipeptide than on the neurtal dipeptide. Moreover, when Bz-Gly-Lys is the substrate, cyclohexanol activates its hydrolysis by CPB by increasing both the enzyme-substrate binding affinity and the rate of the catalytic step, an effect different from that observed when Bz-Gly is the modifier. The anomalous kinetic behavior of CPB is remarkably similar to that of carboxypeptidase A, and is a good indication that both enzymes have very similar structures in and around their respective active sites. A binding site for activator molecules down the cleft of the active site is proposed for CPB to explain the observed kinetic behavior.  相似文献   

12.
The increase in intracellular cyclic GMP concentrations in response to muscarinic-receptor activation in N1E-115 neuroblastoma cells is dependent on extracellular Ca2+ ion. The calcium ionophore A23187 can also evoke an increase in cyclic GMP in the presence of Ca2+ ion. Most (about 85%) of the guanylate cyclase activity of broken-cell preparations is found in the soluble fraction. The soluble enzyme can utilize MnGTP (Km = 55 micrometer), MgGTP (Km = 310 micrometer) and CaGTP (Km greater than 500 micrometer) as substrates. Free GTP is a strong competitive inhibitor (Ki approximately 20 micrometer). The enzyme possesses an allosteric binding site for free metal ions (Ca2+, Mg2+ and Mn2+). The membrane-bound guanylate cyclase is qualitatively similar to the soluble form, but has lower affinity for the metal-GTP substrates. Entry of Ca2+ into cells may increase cyclic GMP concentration by activating guanylate cyclase through an indirect mechanism.  相似文献   

13.
M E Bayliss  J M Prescott 《Biochemistry》1986,25(24):8113-8117
Aeromonas aminopeptidase contains two nonidentical metal binding sites that have been shown by both spectroscopy and kinetics to be capable of interacting with one another [Prescott, J.M., Wagner, F.W., Holmquist, B., & Vallee, B.L. (1985) Biochemistry 24, 5350-5356]. The effects of metal ion substitutions on the susceptibility of the p-nitroanilides of L-alanine, L-valine, and L-leucine--substrates that are hydrolyzed at widely differing rates by native Aeromonas aminopeptidase--were studied by determining values of kcat and Km for the 16 metalloenzymes that result from all possible combinations of Zn2+, Co2+, Ni2+, and Cu2+ in each of the two sites. The different combinations of metal ions and substrates yield a broad range in kinetic values; kcat varies by more than 1800-fold, Km by 3000-fold, and kcat/Km ratios by more than 10,000. L-Leucine-p-nitroanilide is by far the most susceptible of the three substrates, and the hyperactivation previously observed with aminopeptidase containing either Ni2+ or Cu2+ in the first binding site and Zn2+ in the second site occurs only with the two poorer substrates, L-alanine-p-nitroanilide and L-valine-p-nitroanilide. Although the enzyme with Zn2+ in both sites hydrolyzes the substrates with N-terminal alanine and valine poorly, it is extremely effective toward L-leucine-p-nitroanilide. Neither metal binding site can be identified as controlling either Km or kcat; both parameters are influenced by the identity of the metal ions, by the site each occupies, and, most strongly, by the substrate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The dysthrombin, Thrombin Quick, is chromatographically separable into two components designated Thrombin Quick I and Thrombin Quick II. Thrombin Quick II lacks observable catalytic activity toward thrombin substrates. The steady-state kinetics of hydrolysis of benzoylarginine ethyl ester and Tos-Gly-Pro-Arg-p-nitroanilide by Thrombin Quick I are equivalent to those of thrombin. These results, in addition to binding studies with the active site titrant N2-(5-dimethylaminonaphthalene-1-sulfonyl)arginine N-(3-ethyl-1,5-pentanediyl)amide, indicate that binding interactions at the catalytic site of Thrombin Quick I are unaltered. Thrombin Quick I is inhibited by anti-thrombin III at the same rate as thrombin. Steady-state kinetic parameters for the release of fibrinopeptide A indicate defects in both kcat and Km for Thrombin Quick I with kcat/Km equal to 0.012 of the value for thrombin, corresponding to the relative fibrinogen clotting activity of 0.013. The results are interpreted as indicating a defect in Thrombin Quick I at a binding site, external to the catalytic site, which is essential for determining specificity toward fibrinogen. The defect in kcat may result secondarily from small perturbations in the steric relationship of the catalytic triad residues. The rate of hydrolysis by Thrombin Quick I of the protein substrates bovine prothrombin and bovine protein C (in the absence of cofactors) is about one-third of that observed for thrombin, indicating that hydrolysis of these substrates by thrombin involves different specificity determinants than does the hydrolysis of fibrinogen.  相似文献   

15.
The eukaryotic 20 S proteasome contains the following 6 active sites: 2 chymotrypsin-like, 2 trypsin-like, and 2 caspase-like. We previously showed that hydrophobic peptide substrates of the chymotrypsin-like sites allosterically stimulate peptide hydrolysis by the caspase-like sites and their own cleavage. More thorough analysis revealed that these peptides also stimulate peptide hydrolysis by the trypsin-like site. This general activation by hydrophobic peptides occurred even if the chymotrypsin-like sites were occupied by a covalent inhibitor and was highly cooperative, with an average Hill coefficient of 7. Therefore, this stimulation of peptide hydrolysis at all active sites occurs upon binding of hydrophobic peptides to several non-catalytic sites. The stimulation by hydrophobic peptides was not observed in the yeast Delta N alpha 3 mutant 20 S proteasomes, in 20 S-PA26 complexes, or SDS-activated proteasomes and was significantly lower in 26 S proteasomes, all of which appear to have the gated channel in the alpha-rings in an open conformation and hydrolyze peptides at much faster rates than 20 S proteasomes. Also the hydrophobic peptides altered K(m), V(max) of active sites in a similar fashion as PA26 and the Delta N alpha 3 mutation. The activation by hydrophobic peptides was decreased in K(+)-containing buffer, which favors the closed state of the channels. Therefore, hydrophobic peptides stimulate peptide hydrolysis most likely by promoting the opening of the channels in the alpha-rings. During protein breakdown, this peptide-induced channel opening may function to facilitate the release of products from the proteasome.  相似文献   

16.
L J McDonald  M D Mamrack 《Biochemistry》1989,28(26):9926-9932
Bovine heart contains multiple phosphoinositide-specific phospholipase C (PIC) activities separable by ion-exchange chromatography. One PIC activity was purified to apparent homogeneity and migrated as a single band of Mr 85,000 on SDS-PAGE. The purified PIC was characterized with sonicated suspensions of either pure phosphatidylinositol 4,5-bisphosphate (PIP2) or phosphatidylinositol (PI) as substrates. At pH 7, apparent Vmax and Km values were higher for PIP2 than for PI, but the value of Vmax/Km was similar for the two substrates. PIC required Ca2+ for the hydrolysis of either PI or PIP2, and increasing free Ca2+ concentrations from 20 to 300 nM saturated PIC activity. The requirement of Ca2+ for PIC activity and the sensitivity of PIC to Ca2+ concentrations in the physiological range suggested the ion may be a cofactor. The PIC reaction mechanism was determined by two-substrate kinetic analysis; the data fit a model in which PIC contained single sites for Ca2+ and phosphoinositide, and utilized a rapid-equilibrium, random-order ternary mechanism for phosphoinositide hydrolysis. The KCa value for either PI or PIP2 hydrolysis was approximately 30 nM, suggesting resting intracellular free Ca2+ concentrations are sufficient to saturate the Ca2+ site of PIC. La3+ was used as a calcium analogue to modulate PIC activity. Low concentrations of LaCl3 (0.01-0.3 microM) inhibited PIC activity competitively with respect to calcium, consistent with a Ca2+ binding site on the enzyme.  相似文献   

17.
Summary In this paper it is shown that the postsynaptic GABA-receptor chloride ion channel complex is composed of several functional subunits. There are probably at least two stereospecific locations on the receptor for GABA-binding and both must be occupied to obtain an increase in chloride conductance. The interaction between these sites is uncertain but there could be either positive cooperativity between the sites or only a requirement that both sites are occupied without occupation of either site affecting the affinity for GABA of the other site. There is a chloride conductance channel coupled to the GABA receptor which opens for an average of 20 msec and has an average conductance of 18 pS. The GABA-coupled chloride channel may or may not have the same composition as the glycine coupled chloride channel.In addition to the GABA-recognition site and the chloride ion channel, GABA-receptors must have additional binding sites or modulator sites where drugs can bind to modify GABA activation of the GABA-receptor. The convulsant PICRO binds to a site which is independent of the GABA-recognition site and PICRO reduces GABA responses. Barbiturates and benzodiazepines augment GABA-responses without reducing GABA-binding and thus they must bind to a modulator site independent of the GABA recognition site. Whether or not this is the same site as the PICRO binding site is uncertain. Thus, the GABA-receptorchloride ion channel complex is composed of at least: 1) two GABA-binding sites; 2) a chloride ion channel; 3) a convulsant binding site (PICRO-binding site) and 4) an anticonvulsant binding site. This organization serves several obvious purposes. First, since two GABA-molecules are required to activate GABA-coupled chloride ion channels, the dose-response relationship for GABA is sigmoidal and steep. Thus minor shifts in GABA affinity will produce large alterations in GABA-responses and the GABA receptor can be easily modulated. Second, since the receptor has binding sites for convulsant and anticonvulsant compounds which decrease and increase GABA-responses, GABAergic inhibition can easily be modulated.  相似文献   

18.
Bleomycin hydrolase, which hydrolyzes the carboxamide bond in the pyrimidoblamic acid moiety of the bleomycin molecule, also cleaved several p-nitroanilide substrates with a neutral or basic amino acid residue and dipeptide substrates such as L-leucyl-glycine. The activity of bleomycin hydrolase was inhibited by two thiol protease inhibitors, E-64 and leupeptin, as well as by N-ethylmaleimide. These results suggest that bleomycin hydrolase is a thiol aminopeptidase. Magnesium ion, sodium chloride, ethylenediaminetetraacetic acid and 1,2-dihydroxybenzene-3,5-disulfonic acid specifically activated the enzymatic hydrolysis of L-arginine-p-nitroanilide, but did not that of L-leucine-p-nitroanilide. Lineweaver-Burk plots showed that Km values of the enzymatic activity for L-arginine-p-nitroanilide were altered by these reagents, although Vmax values were almost unaltered.  相似文献   

19.
The activation of lysine epsilon-dehydrogenase [EC 1.4.1.] by L-lysine was dependent on lysine concentration and was accompanied by association of the dimeric enzymes to a tetramer. The lysine concentration required for the half-maximal activation was 0.28 mM, which was lower than the Km value for L-lysine. In addition to L-lysine, several compounds, which were neither substrates nor inhibitors, activated the enzyme. The compounds which activated the enzyme have common structural characteristics: they have both a carboxyl group and a hydrophobic side chain. These activators also induced the association of the enzyme. The activation of the enzyme occurred well over the pH range 5.0 to 7.5, and the maximal activation was obtained by preincubation for 5 min at 30 degrees C and pH 7.4, when 5 mM L-lysine or 6-aminocaproate was used as an activator. NADH binding experiments indicated that about 2 mol of NADH bind to 1 mol of the tetrameric enzyme: the dimeric enzyme has one catalytic site. Binding experiments with n-[1-14C]heptanoate and L-[U-14C]lysine showed that approximately 2 mol of ligands bind to 1 mol of the dimeric enzyme and L-lysine could not bind to the catalytic site of the enzyme in the absence of NAD+. These results indicate the presence of one catalytic site and two activator binding binding sites in the dimeric enzyme.  相似文献   

20.
Thiolases are CoA-dependent enzymes which catalyze the formation of a carbon-carbon bond in a Claisen condensation step and its reverse reaction via a thiolytic degradation mechanism. Mitochondrial acetoacetyl-coenzyme A (CoA) thiolase (T2) is important in the pathways for the synthesis and degradation of ketone bodies as well as for the degradation of 2-methylacetoacetyl-CoA. Human T2 deficiency has been identified in more than 60 patients. A unique property of T2 is its activation by potassium ions. High-resolution human T2 crystal structures are reported for the apo form and the CoA complex, with and without a bound potassium ion. The potassium ion is bound near the CoA binding site and the catalytic site. Binding of the potassium ion at this low-affinity binding site causes the rigidification of a CoA binding loop and an active site loop. Unexpectedly, a high-affinity binding site for a chloride ion has also been identified. The chloride ion is copurified, and its binding site is at the dimer interface, near two catalytic loops. A unique property of T2 is its ability to use 2-methyl-branched acetoacetyl-CoA as a substrate, whereas the other structurally characterized thiolases cannot utilize the 2-methylated compounds. The kinetic measurements show that T2 can degrade acetoacetyl-CoA and 2-methylacetoacetyl-CoA with similar catalytic efficiencies. For both substrates, the turnover numbers increase approximately 3-fold when the potassium ion concentration is increased from 0 to 40 mM KCl. The structural analysis of the active site of T2 indicates that the Phe325-Pro326 dipeptide near the catalytic cavity is responsible for the exclusive 2-methyl-branched substrate specificity.  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