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.
Transient kinetic data of ATP binding and cleavage by cardiac myosin subfragment 1 (S1) were obtained by fluorescence stopped flow and analyzed by using computer modeling based on a consecutive, reversible two-step mechanism: (formula: see text) where M1 and M12 denote myosin species with enhanced fluorescence and K'O = K0/(K0[ATP] + 1). The kinetic constants K0, k12, k23, and k32 and the fractional contributions of M1 and M12 to the total fluorescence are analyzed over a range of systematically varied solution parameters. The initial ATP binding equilibrium (K0), which decreases with increasing pH, is facilitated by a positively charged protein residue with a pK of 7.1. An active-site charge of +1.5 is determined from the ionic strength dependence. The rate constants k12, k23, and k32 also exhibit pK's near neutrality but increase with increasing pH. The majority of the large (-54 kJ/mol) negative free energy of ATP binding occurs upon S1 isomerization, k12, and a large increase in entropy (183 J/kmol at 15 degrees C) is associated with the cleavage step. The equilibrium constant for the cleavage step, K2, is determined as 3.5 at pH 7.0, 15 degrees C, and 200 mM ionic strength. There are no significant changes in fractional contributions to total fluorescence enhancement due to solvent-dependent conformational changes of S1 in these data. When values for the combined rate constants are calculated and compared with those determined by graphical analysis, it is observed that graphical analysis overestimates the binding rate constant (K0k12) by 25% and the hydrolysis rate constant (k23 + k32) by as much as 30%.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The photophysics of the complex forming reaction of Ca2+ and Fura-2 are investigated using steady-state and time-resolved fluorescence measurements. The fluorescence decay traces were analyzed with global compartmental analysis yielding the following values for the rate constants at room temperature in aqueous solution with BAPTA as Ca2+ buffer: k01 = 1.2 x 10(9)s-1, k21 = 1.0 x 10(11) M-1 s-1, k02 = 5.5 x 10(8) s-1, k12 = 2.2 x 10(7) s-1, and with EGTA as Ca2+ buffer: k01 = 1.4 x 10(9) s-1, k21 = 5.0 x 10(10) M-1 s-1, k02 = 5.5 x 10(8) s-1, k12 = 3.2 x 10(7) s-1. k01 and k02 denote the respective deactivation rate constants of the Ca2+ free and bound forms of Fura-2 in the excited state. k21 represents the second-order rate constant of binding of Ca2+ and Fura-2 in the excited state, whereas k12 is the first-order rate constant of dissociation of the excited Ca2+:Fura-2 complex. The ionic strength of the solution was shown not to influence the recovered values of the rate constants. From the estimated values of k12 and k21, the dissociation constant K*d in the excited state was calculated. It was found that in EGTA Ca2+ buffer pK*d (3.2) is smaller than pKd (6.9) and that there is negligible interference of the excited-state reaction with the determination of Kd and [Ca2+] from fluorimetric titration curves. Hence, Fura-2 can be safely used as an Ca2+ indicator. From the obtained fluorescence decay parameters and the steady-state excitation spectra, the species-associated excitation spectra of the Ca2+ free and bound forms of Fura-2 were calculated at intermediate Ca2+ concentrations.  相似文献   

3.
Ca2+-dependent proteases isolated from chicken gizzard and bovine aortic smooth muscle were compared with respect to subunit autolysis and the role of autolysis in modulating enzyme activity. The protease isolated from chicken gizzard was a heterodimer consisting of 80,000- and 30,000-dalton subunits. The protease isolated under identical conditions from bovine aorta consisted of 75,000- and 30,000-dalton subunits. In the presence of Ca2+, both enzymes underwent autolysis of their 30,000-dalton subunits with conversion to an 18,000-dalton species. In addition, the 80,000-dalton subunit of the gizzard protease was degraded to a 76,000-dalton form. The Ca2+ concentrations required for autolysis of the 30,000-dalton subunits were different for the two enzymes (i.e. gizzard: K0.5 Ca2+ = 335 microM; aortic: K0.5 Ca2+ = 1,250 microM) although in both cases, stimulation of autolysis by Ca2+ exhibited positive cooperativity. When compared with respect to kinetics of substrate degradation, the native forms of the smooth muscle Ca2+-dependent proteases (gizzard, GIIa = 80,000/30,000-dalton heterodimer; bovine aortic, IIa = 75,000/30,000-dalton heterodimer) exhibited a lag phase in product appearance. On the other hand, the autolyzed forms (gizzard, GIIb = 76,000/18,000-dalton heterodimer; bovine aortic, IIb = 75,000/18,000-dalton heterodimer) exhibited linear rates of substrate degradation. These results were analyzed in terms of autolysis of the 30,000-dalton subunits as determined by the conversion of this subunit to its 18,000 dalton form. For both enzymes, the time course for the autolytic transition, 30,000----18,000 daltons, and Ca2+-dependence of the apparent rate constants for this transition were found to correlate well with the lag phase in enzymatic activity. No such correlation could be established for the 80,000----76,000 dalton autolytic transition of the high molecular mass subunit of the gizzard protease. Our results suggest that catalytic activity of the Ca2+-dependent proteases isolated from gizzard and bovine aortic smooth muscle requires autolysis of the 30,000-dalton subunit. The native or unautolyzed forms of these enzymes appear to be proenzymes that can be activated by autolysis.  相似文献   

4.
Cellobiase has been isolated from the crude cellulase mixture of enzymes of Trichoderma viride using column chromatographic and ion-exchange methods. The steady-state kinetics of the hydrolysis of cellobiose have been investigated as a function of cellobiose and glucose concentrations, pH of the solution, temperature, and dielectric constant, using isopropanol-buffer mixtures. The results show that (i) there is a marked activation of the reaction by initial glucose concentrations of 4 X 10(-3) M to 9 X 10(-2) M and strong inhibition of the reaction at higher initial concentrations, (ii) the log rate -pH curve has a maximum at pH 5.2 and enzyme pK values of 3.5 and 6.8, (iii) the energy of activation at pH 5.1 is 10.2 kcal mol-1 over the temperature range 5-56 degrees C, and (iv) the rate decreases from 0 to 20% (v/v) isopropanol. The hydrolysis by cellobiase (EC 3.2.1.21) of p-nitrophenyl-beta-D-glucoside was examined by pre-steady-state methods in which [enzyme]0 greater than [substrate]0, and by steady-state methods as a function of pH and temperature. The results show (i) a value for k2 of 21 S-1 at pH 7.0 (where k2 is the rate constant for the second step in the assumed two-intermediate mechanism (formula: see text), (ii) a log rate -pH curve, significantly different from that for hydrolysis of cellobiose, in which the rate increases with decreasing pH below pH 4.5, is constant in the region pH 4.5-6, and decreases above pH 6 (exhibiting an enzyme pK value of 7.3), and (iii) an activation energy of 12.5 kcal mol-1 at pH 5.7 over the temperature range 10-60 degrees C.  相似文献   

5.
Ribulose-1,5-bisphosphate carboxylase was activated by incubation with CO2 and Mg2++, and inactivated upon removal of CO2 and Mg2+ by gel filtration. The activation process involved CO2 rather than HCO3-. The activity of the enzyme was dependent upon the preincubation concentrations of CO2 and Mg2+ and upon the preincubation pH, indicating that activation involved the reversible formation of an equilibrium complex of enzyme-CO2-Mg. The initial rate of activation was linearly dependent upon the CO2 concentration but independent of the Mg2+ concentration. Kinetic analyses indicated that the enzyme reacted first with CO2 in a rate-determining and reversible step, followed by a rapid reaction with Mg2+ to form an active ternary complex (see eq 1 in text). The pseudo-first order rate constant, kobsd, for the activation process at constant pH was derived: kobsd=k1[CO2] + (k2k4/k3[Mg2+]). Experimentally, kobsd was shown to be linearly dependent upon the CO2 concentration and inversely dependent upon the Mg2+ concentration. The activity of the enzyme after preincubation to equilibrium at constant concentrations of CO2 and Mg2+ increased as the preincubation pH was raised, indicating that CO2 reacted with an enzyme group whose pK was distinctly alkaline. It is proposed that the activation of ribulose-1, 5-biphosphate carboxylane involves the formation of a carbamate.  相似文献   

6.
Wang H  Vath GM  Gleason KJ  Hanna PE  Wagner CR 《Biochemistry》2004,43(25):8234-8246
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from acetyl coenzyme A (AcCoA) to arylamines, hydrazines, and their N-hydroxylated arylamine metabolites. The recently determined three-dimensional structures of prokaryotic NATs have revealed a cysteine protease-like Cys-His-Asp catalytic triad, which resides in a deep and hydrophobic pocket. This catalytic triad is strictly conserved across all known NATs, including hamster NAT2 (Cys-68, His-107, and Asp-122). Treatment of NAT2 with either iodoacetamide (IAM) or bromoacetamide (BAM) at neutral pH rapidly inactivated the enzyme with second-order rate constants of 802.7 +/- 4.0 and 426.9 +/- 21.0 M(-1) s(-1), respectively. MALDI-TOF and ESI mass spectral analysis established that Cys-68 is the only site of alkylation by IAM. Unlike the case for cysteine proteases, no significant inactivation was observed with either iodoacetic acid (IAA) or bromoacetic acid (BAA). Pre-steady state and steady state kinetic analysis with p-nitrophenyl acetate (PNPA) and NAT2 revealed a single-exponential curve for the acetylation step with a second-order rate constant of (1.4 +/- 0.05) x 10(5) M(-1) s(-1), followed by a slow linear rate of (7.85 +/- 0.65) x 10(-3) s(-1) for the deacetylation step. Studies of the pH dependence of the rate of inactivation with IAM and the rate of acetylation with PNPA revealed similar pK(a)(1) values of 5.23 +/- 0.09 and 5.16 +/- 0.04, respectively, and pK(a)(2) values of 6.95 +/- 0.27 and 6.79 +/- 0.25, respectively. Both rates reached their maximum values at pH 6.4 and decreased by only 30% at pH 9.0. Kinetic studies in the presence of D(2)O revealed a large inverse solvent isotope effect on both inactivation and acetylation of NAT2 [k(H)(inact)/k(D)(inact) = 0.65 +/- 0.02 and (k(2)/K(m)(acetyl))(H)/(k(2)/K(m)(acetyl))(D) = 0.60 +/- 0.03], which were found to be identical to the fractionation factors (Phi) derived from proton inventory studies of the rate of acetylation at pL 6.4 and 8.0. Substitution of the catalytic triad Asp-122 with either alanine or asparagine resulted in the complete loss of protein structural integrity and catalytic activity. From these results, it can be concluded that the catalytic mechanism of NAT2 depends on the formation of a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)). However, in contrast to the case with cysteine proteases, a pH-dependent protein conformational change is likely responsible for the second pK(a), and not deprotonation of the thiolate-imidazolium ion. In addition, substitutions of the triad aspartate are not tolerated. The enzyme appears, therefore, to be engineered to rapidly form a stable acetylated species poised to react with an arylamine substrate.  相似文献   

7.
Alcohol dehydrogenase was purified in 14 h from male Fischer-344 rat livers by differential centrifugation, (NH4)2SO4 precipitation, and chromatography over DEAE-Affi-Gel Blue, Affi-Gel Blue, and AMP-agarose. Following HPLC more than 240-fold purification was obtained. Under denaturing conditions, the enzyme migrated as a single protein band (Mr congruent to 40,000) on 10% sodium dodecyl sulfate-polyacrylamide gels. Under nondenaturing conditions, the protein eluted from an HPLC I-125 column as a symmetrical peak with a constant enzyme specific activity. When examined by analytical isoelectric focusing, two protein and two enzyme activity bands comigrated closely together (broad band) between pH 8.8 and 8.9. The pure enzyme showed pH optima for activity between 8.3 and 8.8 in buffers of 0.5 M Tris-HCl, 50 mM 2-(N-cyclohexylamino)ethanesulfonic acid (CHES), and 50 mM 3-(cyclohexylamino)-1-propanesulfonic acid (CAPS), and above pH 9.0 in 50 mM glycyl-glycine. Kinetic studies with the pure enzyme, in 0.5 M Tris-HCl under varying pH conditions, revealed three characteristic ionization constants for activity: 7.4 (pK1); 8.0-8.1 (pK2), and 9.1 (pK3). The latter two probably represent functional groups in the free enzyme; pK1 may represent a functional group in the enzyme-NAD+ complex. Pure enzyme also was used to determine kinetic constants at 37 degrees C in 0.5 M Tris-HCl buffer, pH 7.4 (I = 0.2). The values obtained were Vmax = 2.21 microM/min/mg enzyme, Km for ethanol = 0.156 mM, Km for NAD+ = 0.176 mM, and a dissociation constant for NAD+ = 0.306 mM. These values were used to extrapolate the forward rate of ethanol oxidation by alcohol dehydrogenase in vivo. At pH 7.4 and 10 mM ethanol, the rate was calculated to be 2.4 microM/min/g liver.  相似文献   

8.
The pH dependence of Vmax and Vmax/Km for hydrolysis of Dnp-Pro-Leu-Gly-Leu-Trp-Ala-D-Arg-NH2 at the Gly-Leu bond by porcine synovial collagenase and gelatinase was determined in the pH range 5-10. Both enzymes exhibited bell-shaped dependencies on pH for these two kinetic parameters, indicating that activity is dependent on at least two ionizable groups, one of which must be unprotonated and the other protonated. For collagenase, Vmax/Km data indicate that in the substrate-free enzyme, these groups have apparent pK values of 7.0 and 9.5, while the Vmax profile indicates similar pK values of 6.8 and 10.1 for the enzyme-substrate complex. The corresponding pH profiles of gelatinase were similar to those of collagenase, indicating the importance of groups with apparent pK values of 5.9 and 10.0 for the free enzyme and 5.9 and 11.1 for the enzyme-substrate complex. When these kinetic constants were determined in D2O using the peptide substrate, there was no significant effect on Vmax or Km for collagenase or Km for gelatinase. However, there was a deuterium isotope effect of approximately 1.5 on Vmax for gelatinase. These results indicate that a proton transfer step is not involved in the rate-limiting step for collagenase, but may be limiting with gelatinase. The Arrhenius activation energies for peptide bond hydrolysis of the synthetic peptide as well as the natural substrates were also determined for both enzymes. The activation energy (81 kcal) for hydrolysis of collagen by collagenase was nine times greater than that determined for the synthetic substrate (9.2 kcal). In contrast, the activation energy for hydrolysis of gelatin by gelatinase (26.3 kcal) was only 2.4 times greater than that for the synthetic substrate (11 kcal).  相似文献   

9.
Adachi MS  Torres JM  Fitzpatrick PF 《Biochemistry》2010,49(49):10440-10448
The flavoprotein oxidase Fms1 from Saccharomyces cerevisiae catalyzes the oxidation of spermine and N(1)-acetylspermine to yield spermidine and 3-aminopropanal or N-acetyl-3-aminopropanal. The kinetic mechanism of the enzyme has been determined with both substrates. The initial velocity patterns are ping-pong, consistent with reduction being kinetically irreversible. Reduction of Fms1 by either substrate is biphasic. The rate constant for the rapid phase varies with the substrate concentration, with limiting rates for reduction of the enzyme of 126 and 1410 s(-1) and apparent K(d) values of 24.3 and 484 μM for spermine and N(1)-acetylspermine, respectively. The rapid phase is followed by a concentration-independent phase that is slower than turnover. The reaction of the reduced enzyme with oxygen is monophasic, with a rate constant of 402 mM(-1) s(-1) with spermine at 25 °C and 204 mM(-1) s(-1) with N(1)-acetylspermine at 4 °C and pH 9.0. This step is followed by rate-limiting product dissociation. The k(cat)/K(amine)-pH profiles are bell-shaped, with an average pK(a) value of 9.3 with spermine and pK(a) values of 8.3 and 9.6 with N(1)-acetylspermine. Both profiles are consistent with the active forms of substrates having two charged nitrogens. The pH profiles for the rate constant for flavin reduction show pK(a) values of 8.3 and 7.2 for spermine and N(1)-acetylspermine, respectively, for groups that must be unprotonated; these pK(a) values are assigned to the substrate N4. The k(cat)/K(O(2))-pH profiles show pK(a) values of 7.5 for spermine and 6.8 for N(1)-acetylspermine. With both substrates, the k(cat) value decreases when a single residue is protonated.  相似文献   

10.
N Ahn  J P Klinman 《Biochemistry》1983,22(13):3096-3106
Dopamine beta-monooxygenase catalyzes a reaction in which 2 mol of protons are consumed for each turnover of substrate. Studies of the pH dependence of initial rate parameters (Vmax and Vmax/Km) and their primary hydrogen isotope effects show that at least two ionizable residues are involved in catalysis. One residue (B1, pK = 5.6-5.8) must be protonated prior to the carbon-hydrogen bond cleavage step, implying a role for general-acid catalysis in substrate activation. A second protonated residue (B2, pK = 5.2-5.4) facilitates, but is not required for, product release. Recent measurement of the intrinsic isotope effect for dopamine beta-monoxygenase [Miller, S. M., & Klinman, J. P. (1983) Biochemistry (preceding paper in this issue)] allows an analysis of the pH dependence of rate constant ratios and in selected instances individual rate constants. We demonstrate large changes in the rate-determining step as well as an unprecedented inversion in the kinetic order of substrate release from ternary complex over an interval of 2 pH units. Previously, fumarate has been used in dopamine beta-monooxygenase assays because of its property of enzyme activation. Studies of the pH behavior in the presence of saturating concentrations of fumarate have shown two causes of the activation: (i) fumarate perturbs the pK of B1 to pK = 6.6-6.8 such that the residue remains protonated and the enzyme optimally active over a wider pH range; (ii) fumarate decreases the rate of dopamine release from the ternary enzyme-substrate complex, increasing the equilibrium association constant for dopamine binding. Both effects are consistent with a simple electrostatic stabilization of bound cationic charges by the dianionic form of fumarate.  相似文献   

11.
Recent studies on calcium-activated protease (CAF) have indicated that there are two forms of this enzyme, one requiring millimolar levels of Ca2+ and one requiring micromolar levels of Ca2+ for maximal activation. We have attempted to elucidate the biochemical nature of the difference between the two forms by the use of one dimensional peptide maps and immunoautoradiography, and have found that the 80,000-dalton subunits from the two forms differ substantially while the 30,000-dalton subunit appear to be identical.  相似文献   

12.
The fluorescent reagent, S-mercuric N-dansyl-cysteine, reacts specifically with thiols of the purified Ca2+-ATPase of the sarcoplasmic reticulum, producing an increase of fluorescence of fluorescence intensity at 500 nm (lambda ex = 335 nm). The reaction is stoichiometric, and the increase of the fluorescence intensity is proportional to the number of blocked thiols. Twelve reactive thiols per 10(5) daltons of ATPase peptide fall into roughly three classes. Blocking of the most reactive thiol entails little inhibition of enzyme activity. Blocking of the five thiols reacting next (intermediate class) results in almost complete inhibition of both phosphorylated intermediate formation and ATP hydrolysis. The second order rate constants of the reaction of thiols have been determined by stopped flow studies. The most reactive thiol and the six least reactive thiols can each be treated as a single class with respect to the rate constant; five thiols of intermediate reactivity appear to have different rate constants (k2, k3, ..k6). Of these constants, k1, corresponding to the most reactive thiol, does not change with [Ca2+]. Upon increasing [Ca2+] from 10(-9) to 10(-5) M, k2 increase and k7-12 decreases; the changes roughly parallel the activation of ATPase activity and the Ca2+ binding to the high affinity alpha sites (Ikemoto, N. (1975) J. Biol. Chem. 250, 7219-7224). Upon further increase of [Ca2+] k2 decreases and k7-12 increase, in parallel with the inhibition of ATPase activity and with the Ca2+ binding to the low affinity gamma sites.  相似文献   

13.
The kinetics of binding and dissociation for the progesterone-binding globulin (PBG)-progesterone complex have been measured as a function of pH. The association rate constant appears to be independent of pH from pH to 10 with an average value of kon = 8.5 X 10(7)M-1 S-1. The dissociation rate constant is strongly pH dependent with the dependency defined by: koff = k0 (1 + [H+]/K1 + K2/[H+])(1 + K3*/[H+])/(1 + K3/[H+]). The best values for the various parameters were k0 = 0.0785 s-1, pK1 = 5.30, pK2 = 10.54, pK3* = 7.41, and pK3 = 7.21. Simpler expressions were inadequate to fit the data, and it was concluded that at least three ionizing residues are responsible for the stability of the PBG-progesterone complex. The affinity constant was determined by equilibrium dialysis over the range of pH 3 to 12. The ratio of the association and dissociation rate constants is in agreement with the affinity constant from pH 6.5 to 10.5. The influence of pH on the conformation and binding activity of PBG was also investigated. Denaturation by acid, base, or guanidine hydrochloride leads to a reversible loss of binding activity. Regain of binding activity in all cases is slow with half-times of 0.5 to 2.7 h, depending on conditions. The rate of acid denaturation was found to be incompletely protonated at pH 1.4, suggesting a buried carboxylic acid residue. The slow renaturation of PBG might be due to the difficulty of burying a charged residue in the protein's interior coupled with steric hindrance by the large carbohydrate moiety of PBG.  相似文献   

14.
The actin-myosin lattice spacing of rabbit psoas fibers was osmotically compressed with a dextran T-500, and its effect on the elementary steps of the cross-bridge cycle was investigated. Experiments were performed at the saturating Ca (pCa 4.5-4.9), 200 mM ionic strength, pH 7.0, and at 20 degrees C, and the results were analyzed by the following cross-bridge scheme: [formula: see text] where A = actin, M = myosin head, S = MgATP, D = MgADP, and P = Pi = phosphate. From MgATP and MgADP studies on exponential process (C) and (D), the association constants of cross-bridges to MgADP (K0), MgATP (K1a), the rate constants of the isomerization of the AM S state (k1b and k-1b), and the rate constants of the cross-bridge detachment step (k2 and k-2) were deduced. From Pi study on process (B), the rate constants of the cross-bridge attachment (power stroke) step (k4- and k-4) and the association constant of Pi ions to cross-bridges (K5) were deduced. From ATP hydrolysis measurement, the rate constant of ADP-isomerization (rate-limiting) step (k6) was deduced. These kinetic constants were studied as functions of dextran concentrations. Our results show that nucleotide binding, the ATP-isomerization, and the cross-bridge detachment steps are minimally affected by the compression. The rate constant of the reverse power stroke step (k-4) decreases with mild compression (0-6.3% dextran), presumably because of the stabilization of the attached cross-bridges in the AM*DP state. The rate constant of the power stroke step (k4) does not change with mild compression, but it decreases with higher compression (> 6.3% dextran), presumably because of an increased difficulty in performing the power stroke. These results are consistent with the observation that isometric tension increases with a low level of compression and decreases with a high level of compression. Our results also show that the association constant K5 of Pi with cross-bridge state AM*D is not changed with compression. Our result further show that the ATP hydrolysis rate decreased with compression, and that the rate constants of the ADP-isomerization step (k6) becomes progressively less with compression. The effect of compression on the power stroke step and rate-limiting step implies that a large-scale molecular rearrangement in the myosin head takes place in these two slow reaction steps.  相似文献   

15.
J Kuret  H Schulman 《Biochemistry》1984,23(23):5495-5504
A soluble Ca2+/calmodulin-dependent protein kinase has been purified from rat brain to near homogeneity by using casein as substrate. The enzyme was purified by using hydroxylapatite adsorption chromatography, phosphocellulose ion-exchange chromatography, Sepharose 6B gel filtration, affinity chromatography using calmodulin-Sepharose 4B, and ammonium sulfate precipitation. On sodium dodecyl sulfate (NaDodSO4)-polyacrylamide gels, the purified enzyme consists of three protein bands: a single polypeptide of 51 000 daltons and a doublet of 60 000 daltons. Measurements of the Stokes radius by gel filtration (81.3 +/- 3.7 A) and the sedimentation coefficient by sucrose density sedimentation (13.7 +/- 0.7 S) were used to calculate a native molecular mass of 460 000 +/- 29 000 daltons. The kinase autophosphorylated both the 51 000-dalton polypeptide and the 60 000-dalton doublet, resulting in a decreased mobility in NaDodSO4 gels. Comparison of the phosphopeptides produced by partial proteolysis of autophosphorylated enzyme reveals substantial similarities between subunits. These patterns, however, suggest that the 51 000-dalton subunit is not a proteolytic fragment of the 60 000-dalton doublet. Purified Ca2+/calmodulin-dependent casein kinase activity was dependent upon Ca2+, calmodulin, and ATP X Mg2+ or ATP X Mn2+ when measured under saturating casein concentrations. Co2+, Mn2+, and La3+ could substitute for Ca2+ in the presence of Mg2+ and saturating calmodulin concentrations. In addition to casein, the purified enzyme displayed a broad substrate specificity which suggests that it may be a "general" protein kinase with the potential for mediating numerous processes in brain and possibly other tissues.  相似文献   

16.
A calmodulin-dependent protein phosphatase has been identified in human platelets by its cross-reactivity with an antibody developed against a bovine brain calmodulin-dependent protein phosphatase and by its calmodulin-stimulated dephosphorylation of 32P-labeled substrates. The platelet enzyme was partially purified to separate it from calmodulin and calmodulin-independent phosphatases. The partially purified enzyme was stimulated by calmodulin, requiring 15 nM calmodulin for half-maximal activation. Calmodulin increased the Vmax of the phosphatase, with no significant effect on its Km. The enzyme was stimulated irreversibly and made calmodulin-independent by limited proteolysis. The optimal pH for the phosphatase was 7.5. After partial purification, phosphatase activity was significantly increased in the presence of Mn2+ and Ca2+ over that observed in the presence of Ca2+ alone. The enzyme effectively dephosphorylated casein, histone, protamine, and platelet actin. The holophosphatase was estimated to have a molecular weight of 76,900 as determined by sedimentation on sucrose gradients. Immunoblotting techniques using an antibody against the brain phosphatase suggests that the enzyme consists of 2 subunits of 60,000 and 16,500 daltons; the 60,000-dalton subunit co-migrates in sodium dodecyl sulfate-polyacrylamide gel electrophoresis with a 60,000-dalton calmodulin-binding protein in the platelet suggesting that it is the calmodulin-binding subunit of the enzyme. The identification of a calmodulin-dependent protein phosphatase in human platelets suggests a role for Ca2+-dependent dephosphorylation in platelet activation.  相似文献   

17.
To better understand substrate recognition and catalysis by RNase III, we examined steady-state and pre-steady-state reaction kinetics, and changes in intrinsic enzyme fluorescence. The multiple turnover cleavage of a model RNA substrate shows a pre-steady-state burst of product formation followed by a slower phase, indicating that the steady-state reaction rate is not limited by substrate cleavage. RNase III catalyzed hydrolysis is slower at low pH, permitting the use of pre-steady-state kinetics to measure the dissociation constant for formation of the enzyme-substrate complex (K(d)=5.4(+/-0.6) nM), and the rate constant for phosphodiester bond cleavage (k(c)=1.160(+/-0.001) min(-1), pH 5.4). Isotope incorporation analysis shows that a single solvent oxygen atom is incorporated into the 5' phosphate of the RNA product, which demonstrates that the cleavage step is irreversible. Analysis of the pH dependence of the single turnover rate constant, k(c), fits best to a model for two or more titratable groups with pK(a) of ca 5.6, suggesting a role for conserved acidic residues in catalysis. Additionally, we find that k(c) is dependent on the pK(a) value of the hydrated divalent metal ion included in the reaction, providing evidence for participation of a metal ion hydroxide in catalysis, potentially in developing the nucleophile for the hydrolysis reaction. In order to assess whether conformational changes also contribute to the enzyme mechanism, we monitored intrinsic tryptophan fluorescence. During a single round of binding and cleavage by the enzyme we detect a biphasic change in fluorescence. The rate of the initial increase in fluorescence was dependent on substrate concentration yielding a second-order rate constant of 1.0(+/-0.1)x10(8) M(-1) s(-1), while the rate constant of the second phase was concentration independent (6.4(+/-0.8) s(-1); pH 7.3). These data, together with the unique dependence of each phase on divalent metal ion identity and pH, support the hypothesis that the two fluorescence transitions, which we attribute to conformational changes, correlate with substrate binding and catalysis.  相似文献   

18.
Glutamate dehydrogenase from Candida utilis undergoes a reversible conformational transition between an active and an inactive state at low pH AND low temperature. This conformational transition can also be followed by fluorescence measurements. The temperature-dependent equilibrium between the active and the inactive state is characterized by a transition temperature of 10.7 degrees C and a delta H value of 148 kcal/mol (620 kJ/mol). The temperature dependence of the enzymic activity above 15 degrees C yields an activation energy of 15 kcal/mol (63 kJ/mol), a larger value than that for the beef liver enzyme (9 kcal/mol; 38 kJ/mol). In contrast to the yeast enzyme the Arrhenius plot is linear and, therefore, the beef liver enzyme is not transformed into an inactive conformation at low temperatures. Sedimentation analysis shows that the inactivation of the Candida utilis enzyme is not caused by change in the quaternary structure. The pH dependence of the conformational transition at low pH measured by fluorescence change is characterized by a pK value of 7.01 for the enzyme in the absence and of 6.89 for the enzyme in the presence of 2-oxoglutarate with a Hill coefficient of 3.4 in both cases. Similar results are found when the pH dependence of the enzymic activity is analyzed. With the beef liver enzyme the same pK value is obtained but with a Hill coefficient of 1 indicating cooperativity only in the case of the Candida utilis enzyme. The best fit of the pH dependence of the rate constants of the fluorescence changes was obtained with pK values of 7.45 and 6.45 for the active and the inactive state respectively. In this model the lowest time constant which is obtained at the pH of the equilibrium was found to be 0.05 s-1. Preincubation experiments with the substrate 2-oxoglutarate but not with the coenzyme shift the equilibrium to the active conformation. The coenzyme obviously reduces the rate constant of the conformational transition. The sedimentation coefficient (SO20, w) and the molecular weight were found to be 11.0 S and 276 000, respectively. The enzyme molecule is built up by six polypeptide chains each having a molecular weight of 47 000.  相似文献   

19.
The complete time course of the hydrolysis of p-nitrophenyl phosphate catalyzed by the low molecular weight (acid) phosphotyrosyl protein phosphatase from bovine heart was elucidated and analyzed in detail. Burst titration kinetics were demonstrated for the first time with this class of enzyme. At pH 7.0, 4.5 degrees C, a transient pre-steady-state "burst" of p-nitrophenol was formed with a rate constant of 48 s-1. The burst was effectively stoichiometric and corresponded to a single enzyme active site/molecule. The burst was followed by a slow steady-state turnover of the phosphoenzyme intermediate with a rate constant of 1.2 s-1. Product inhibition studies indicated an ordered uni-bi kinetic scheme for the hydrolysis. Partition experiments conducted for several substrates revealed a constant product ratio. Vmax was constant for these substrates, and the overall rate of hydrolysis was increased greatly in the presence of alcohol acceptors. An enzyme-catalyzed 18O exchange between inorganic phosphate and water was detected and occurred with kcat = 4.47 x 10(-3) s-1 at pH 5.0, 37 degrees C. These results were all consistent with the existence of a phosphoenzyme intermediate in the catalytic pathway and with the breakdown of the intermediate being the rate-limiting step. The true Michaelis binding constant Ks = 6.0 mM, the apparent Km = 0.38 mM, and the rate constants for phosphorylation (k2 = 540 s-1) and dephosphorylation (k3 = 36.5 s-1) were determined under steady-state conditions with p-nitrophenyl phosphate at pH 5.0 and 37 degrees C in the presence of phosphate acceptors. The energies of activation for the enzyme-catalyzed hydrolysis at pH 5.0 and 7.0 were 13.6 and 14.1 kcal/mol, respectively. The activation energy for the enzyme-catalyzed medium 18O exchange between phosphate and water was 20.2 kcal/mol. Using the available equilibrium and rate constants, an energetic diagram was constructed for the enzyme-catalyzed reaction.  相似文献   

20.
V C Sekhar  B V Plapp 《Biochemistry》1988,27(14):5082-5088
The binding of NAD+ to liver alcohol dehydrogenase was studied by stopped-flow techniques in the pH range from 6.1 to 10.9 at 25 degrees C. Varying the concentrations of NAD+ and a substrate analogue used to trap the enzyme-NAD+ complex gave saturation kinetics. The same maximum rate constants were obtained with or without the trapping agent and by following the reaction with protein fluorescence or absorbance of a ternary complex. The data fit a mechanism with diffusion-controlled association of enzyme and NAD+, followed by an isomerization with a forward rate constant of 500 s-1 at pH 8: E E-NAD+ *E-NAD+. The isomerization may be related to the conformational change determined by X-ray crystallography of free enzyme and enzyme-coenzyme complexes. Overall bimolecular rate constants for NAD+ binding show a bell-shaped pH dependence with apparent pK values at 6.9 and 9.0. Acetimidylation of epsilon-amino groups shifts the upper pK to a value of 11 or higher, suggesting that Lys-228 is responsible for the pK of 9.0. Formation of the enzyme-imidazole complex abolishes the pK value of 6.9, suggesting that a hydrogen-bonded system extending from the zinc-bound water to His-51 is responsible for this pK value. The rates of isomerization of E-NAD+ and of pyrazole binding were maximal at pH below a pK of about 8, which is attributable to the hydrogen-bonded system. Acetimidylation of lysines or displacement of zinc-water with imidazole had little effect on the rate of isomerization of the E-NAD+ complex.(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