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 15 毫秒
1.
As a substitute for M(H2O)2+6, Co(NH3)3+6 was found to activate calcineurin with para-nitrophenyl phosphate as substrate. Kinetics for calcineurin catalyzed hydrolysis of para-nitrophenyl phosphate at pH 7.0 with Mn2+, Mg2+, Co2+, and Co(NH3)3+6 were compared. Although kcat and Km were different with the metals, values of kcat/Km were nearly identical for Mn2+ and Mg2+, but lower for Co2+ and Co(NH3)3+6. The concentration of each metal providing half-maximal activation, designated Kact, was evaluated as 15.9 mM for Co(NH3)3+6, compared to Kact = 0.17 mM for Mn2+ and Co2+ and 6.3 mM for Mg2+, respectively. Comparing kcat/Kcat showed that Co(NH3)3+6 was a 170-fold poorer activator of calcineurin than was Mn2+, but only 1.5-fold poorer than Mg2+. Activation by Co(NH3)3+6 indicated that activation of calcineurin by exogenous metal ions can proceed via an outer coordination sphere reaction mechanism with no requirement for the direct coordination of substrate by metal. Because Co(NH3)3+6 was found to support calcineurin activity, the related compound [Co-(ethylenediamine)3]3+ (or Co(en)3+3) was tested as a possible activator. Co(en)3+3 did not support calcineurin activity but did inhibit calcineurin. Co(en)3+3 showed competitive inhibition kinetics with either Mn2+ or pNPP as the varied ligand and the other at a fixed, subsaturating concentration. Inorganic phosphate was used as a known competitive inhibitor to pNPP (B. L. Martin and D. J. Graves, J. Biol. Chem. 261, 14545-14550, 1986) and showed uncompetitive inhibition with Mn2+ as the varied ligand. These patterns are consistent with the mechanism of ligand binding to calcineurin being ordered with metal preceding substrate. Prior formation of a metal-substrate complex was not required for association with calcineurin.  相似文献   

2.
Inhibition of sarcoplasmic reticulum Ca2+-ATPase by Mg2+ at high pH   总被引:1,自引:0,他引:1  
Steady state turnover of Ca2+-ATPase of sarcoplasmic reticulum has generally been reported to have a bell-shaped pH profile, with an optimum near pH 7.0. While a free [Mg2+] of 2 mM is optimal for activity at pH 7.0, it was found that this level was markedly inhibitory (K1/2 = 2 mM) at pH 8.0, thus accounting for the generally observed low activity at high pH. High activity was restored at pH 8.0 using an optimum free [Mg2+] of 0.2 mM. The mechanism of the Mg2+-dependent inhibition at pH 8.0 was probed. Inhibition was not due to Mg2+ competition with Ca2+ for cytoplasmic transport sites nor to inhibition of formation of steady state phosphoenzyme from ATP. Mg2+ inhibited (K1/2 = 1.8 mM) decay of steady state phosphoenzyme; thus, the locus of inhibition was one of the phosphoenzyme interconversion steps. Transient kinetic experiments showed that Mg2+ competitively inhibited (Ki = 0.7 mM) binding of Ca2+ to lumenal transport sites, blocking the ability of Ca2+ to reverse the catalytic cycle to form ADP-sensitive, from ADP-insensitive, phosphoenzyme. The data were consistent with a hypothesis in which Mg2+ binds lumenal Ca2+ transport sites with progressively higher affinity at higher pH to form a dead-end complex; its dissociation would then be rate-limiting during steady state turnover.  相似文献   

3.
A D Hall  A Williams 《Biochemistry》1986,25(17):4784-4790
Values of kcat and Km have been measured for the Escherichia coli alkaline phosphatase catalyzed hydrolysis of 18 aryl and 12 alkyl monophosphate esters at pH 8.00 and 25 degrees C. A Br?nsted plot of log (kcat/Km) (M-1 s-1) vs. the pK of the leaving hydroxyl group exhibits two regression lines: log (kcat/Km) = -0.19 (+/- 0.02) pKArOH + 8.14 (+/- 0.15) log (kcat/Km) = -0.19 (+/- 0.01) pKROH + 5.89 (+/- 0.17) Alkyl phosphates with aryl or large lipophilic side chains are not correlated by the above equations and occupy positions intermediate between the two lines. The observed change in effective charge on the leaving oxygen of the ester (-0.2) is very small, consistent with substantial electrophilic participation of the enzyme with this atom. Cyclohexylammonium ion is a noncompetitive inhibitor against 4-nitrophenyl phosphate substrate at pH 8.00, and neutral phenol is a competitive inhibitor (Ki = 82.6 mM); these data and the 100-fold larger reactivity of aryl over alkyl esters are consistent with the existence of a lipophilic binding site for the leaving group of the substrate. The absence of a major steric effect in kcat/Km for substituted aryl esters confirms that the leaving group in the enzyme--substrate complex points away from the surface of the enzyme. Arguments are advanced to exclude a dissociative mechanism (involving a metaphosphate ion) for the enzyme-catalyzed substitution at phosphorus.  相似文献   

4.
A high basal level of phosphorylation (approx. 70% of the optimal Na+-dependent phosphorylation level) is observed in 50 mM imidazole-HCl (pH 7.0), in the absence of added Na+ and K+ and the presence of 10-100 microM Mg2+. In 50 mM Tris-HCl (pH 7.0) the basal level is only 5%, irrespective of the Mg2+ concentration. Nevertheless, imidazole is a less effective activator of phosphorylation than Na+ (Km imidazole-H+ 5.9 mM, Km Na+ 2 mM under comparable conditions). Imidazole-activated phosphorylation is strongly pH dependent, being optimal at pH less than or equal to 7 and minimal at pH greater than or equal to 8, while Na+-activated phosphorylation is optimal at pH 7.4. This suggests that imidazole-H+ is the activating species. Imidazole facilitates Na+-stimulated phosphorylation. The Km for Na+ decreases from 0.63 mM at 5 mM imidazole-HCl to 0.21 mM at 50 mM imidazole-HCl (pH 7; 0.1 mM Mg2+ in all cases). Imidazole-activated phosphorylation is more sensitive to inhibition by K+ (I50 = 12.5 microM) than Na+-activated phosphorylation (I50 = 180 microM). Mg2+ antagonizes activation by imidazole-H+ and also inhibition by K+. The Ki value for Mg2+ (approx. 0.3 mM) is the same for the two antagonistic effects. Tris buffer (pH 7.0) inhibits imidazole-activated phosphorylation with an I50 value of 30 mM in 50 mM imidazole-HCl (pH 7.0) plus 0.1 mM Mg2+. We conclude that imidazole-H+, but not Tris-H+, can replace Na+ as an activator of ATP-dependent phosphorylation, primarily by shifting the E2----E1 transition to the right, leading to a phosphorylating E1 conformation which is different from that in Tris buffer.  相似文献   

5.
A cDNA clone encoding amorpha-4,11-diene synthase from Artemisia annua was subcloned into a bacterial expression vector in frame with a His6-tag. Recombinant amorpha-4,11-diene synthase was produced in Escherichia coli and purified to apparent homogeneity. The enzyme showed pH optimum at pH 6.5, and a minimum at pH 7.5. Substantial activity was observed in the presence of Mg2+, Mn2+ or Co2+ as cofactor. The enzyme exhibits a low activity in the presence of Ni2+ and essentially no activity with Cu2+ or Zn2+. The sesquiterpenoids produced from farnesyl diphosphate in the presence of Mg2+ were analyzed by GC-MS. In addition to amorpha-4,11-diene, 15 sesquiterpenoids were produced. Only small quantitative differences in product pattern were observed at pH 6.5, 7.5, or 9.5. Amorpha-4,11-diene synthase showed significant increased product selectivity in the presence of Mn2+ or Co2+. Km for farnesyl diphosphate was 3.3, 8.0, and 0.7 microM in the presence of Mg2+, Mn2+ or Co2+, respectively. The corresponding kcat-values were 6.8, 15.0, and 1.3 x 10(-3) s(-1), respectively. Km and kcat for geranyl diphosphate were 16.9 microM and 7.0 x 10(-4) s(-1), respectively, at pH 6.5, in the presence of Mn2+.  相似文献   

6.
We have measured the pH dependence of kcat and kcat/Km for CO2 hydration catalyzed by both native Zn2+-and metallo-substituted Co2+-bovine carbonic anhydrase II in the absence of inhibitory ions. For the Zn2+-enzyme, the pKa values controlling kcat and kcat/Km profiles are similar, but for the Co2+-enzyme the values are about 0.6 pH units apart. Computer simulations of a metal-hydroxide mechanism of carbonic anhydrase suggest that the data for both native and Co2+-carbonic anhydrase can be accounted for by the same mechanism of action, if we postulate that the substitution of Co2+ for Zn2+ in the active site causes a separation of about 0.6 pH units in the pKa values of His-64 and the metal-bound water molecule. We have also measured the activation parameters for kcat and kcat/Km for Co2+-substituted carbonic anhydrase II-catalyzed CO2 hydration and have compared these values to those obtained previously for the native Zn2+-enzyme. For kcat and kcat/Km we obtain an enthalpy of activation of 4.4 +/- 0.6 and approximately 0 kcal mol-1, respectively. The corresponding entropies of activation are -18 +/- 2 and -27 +/- 2 cal mol-1 K-1.  相似文献   

7.
The activity of yeast enolase is inhibited by Li+ and Na+. At pH 7.1, inhibition by Li+ is "mixed" with respect to Mg2+; both Vmax and Km (Mg2+) are increased by Li+. The inhibition by Li+ appears to be partial, indicating that enzyme with Li+ bound is active. The step inhibited by Li+ cannot be proton abstraction since Li+ decreases the kinetic isotope effect on Vmax. At pH 9.2, where proton abstraction is no longer partially rate-limiting, inhibiton by Li+ is competitive with respect to Mg2+. The rate of enzyme-catalyzed exchange of the C-2 hydrogen with solvent is not affected by Li+. We interpret these results as follows: Li+ (and Na+) binds to enolase and decreases the rate of at least one step in the mechanism. At pH 7.1, this step is partially rate-limiting; at pH 9.2, this step is a fast step in the reaction. The step inhibited by Li+ cannot be proton abstraction but may be release of product (phosphoenol pyruvate) or Mg2+.  相似文献   

8.
Z Y Zhang  R L Van Etten 《Biochemistry》1991,30(37):8954-8959
The kcat and Km values for the bovine heart low molecular weight phosphotyrosyl protein phosphatase catalyzed hydrolysis of 16 aryl phosphate monoesters and of five alkyl phosphate monoesters having the structure Ar(CH2)nOPO3H2 (n = 1-5) were measured at pH 5.0 and 37 degrees C. With the exception of alpha-naphthyl phosphate and 2-chlorophenyl phosphate, which are subject to steric effects, the values of kcat are effectively constant for the aryl phosphate monoesters. This is consistent with the catalysis being nucleophilic in nature, with the existence of a common covalent phosphoenzyme intermediate, and with the breakdown of this intermediate being rate-limiting. In contrast, kcat for the alkyl phosphate monoesters is much smaller and the rate-limiting step for these substrates is interpreted to be the phosphorylation of the enzyme. A single linear correlation is observed for a plot of log (kcat/Km) vs leaving group pKa for both classes of substrates at pH 5.0: log (kcat/Km) = -0.28pKa + 6.88 (n = 19, r = 0.89), indicating a uniform catalytic mechanism for the phosphorylation event. The small change in effective charge (-0.28) on the departing oxygen of the substrate is similar to that observed in the specific acid catalyzed hydrolysis of monophosphate monoanions (-0.27) and is consistent with a strong electrophilic interaction of the enzyme with this oxygen atom in the transition state. The D2O solvent isotope effect and proton inventory experiments indicate that only one proton is "in flight" in the transition state of the phosphorylation process and that this proton transfer is responsible for the reduction of effective charge on the leaving oxygen.  相似文献   

9.
McCain DF  Grzyska PK  Wu L  Hengge AC  Zhang ZY 《Biochemistry》2004,43(25):8256-8264
Protein tyrosine phosphatases (PTPs) constitute a large family of signaling enzymes that include both tyrosine specific and dual-specificity phosphatases that hydrolyze pSer/Thr in addition to pTyr. Previous mechanistic studies of PTPs have relied on the highly activated substrate p-nitrophenyl phosphate (pNPP), an aryl phosphate with a leaving group pK(a) of 7. In the study presented here, we employ m-nitrobenzyl phosphate (mNBP), an alkyl phosphate with a leaving group pK(a) of 14.9, which mimics the physiological substrates of the PTPs. We have carried out pH dependence and kinetic isotope effect measurements to characterize the mechanism of two important members of the PTP superfamily: Yersinia PTP (YopH) and Cdc25A. Both YopH and Cdc25A exhibit bell-shaped pH-rate profiles for the hydrolysis of mNBP, consistent with general acid catalysis. The slightly inverse (18)(V/K)(nonbridge) isotope effects (0.9999 for YopH and 0.9983 for Cdc25A) indicate a loose transition state with little nucleophilic participation for both enzymes. The smaller (18)(V/K)(bridge) primary isotope effects (0.9995 for YopH and 1.0012 for Cdc25A) relative to the corresponding isotope effects for pNPP hydrolysis suggest that protonation of the leaving group oxygen at the transition state by the general acid is ahead of P-O bond fission with the alkyl substrate, while general acid catalysis of pNPP by YopH is more synchronous with P-O bond fission. The isotope effect data also confirm findings from previous studies that Cdc25A utilizes general acid catalysis for substrates with a leaving group pK(a) of >8, but not for pNPP. Interestingly, the difference in the kinetic isotope effects for the reactions of aryl phosphate pNPP and alkyl phosphate mNBP by the PTPs parallels what is observed in the uncatalyzed reactions of their monoanions. In these reactions, the leaving group is protonated in the transition state, as is the case in PTP-catalyzed reactions. Also, the phosphoryl group in the transition states of the enzymatic reactions does not differ substantially from those of the uncatalyzed reactions. These results provide further evidence that these enzymes do not change the transition state but simply stabilize it.  相似文献   

10.
Markham GD  Bock CL  Schalk-Hihi C 《Biochemistry》1999,38(14):4433-4440
Inosine-5'-monophosphate dehydrogenase (IMPDH) catalyzes the K+-dependent reaction IMP + NAD + H2O --> XMP + NADH + H+ which is the rate-limiting step in guanine nucleotide biosynthesis. The catalytic mechanism of the human type-II IMPDH isozyme has been studied by measurement of the pH dependencies of the normal reaction, of the hydrolysis of 2-chloro-IMP (which yields XMP and Cl- in the absence of NAD), and of inactivation by the affinity label 6-chloro-purine-ribotide (6-Cl-PRT). The pH dependence of the IMPDH reaction shows bell-shaped profiles for kcat and the kcat/Km values for both IMP and NAD, illustrating the involvement of both acidic and basic groups in catalysis. Half-maximal kcat values occur at pH values of 7.2 and 9.8; similar pK values of 6.9 and 9.4 are seen in the kcat/Km profile for NAD. The kcat/Km profile for IMP, which binds first in the predominantly ordered kinetic mechanism, shows pK values of 8.1 and 7.3 for acidic and basic groups, respectively. None of the kinetic pK values correspond to ionizations of the free substrates and thus reflect ionization of the enzyme or enzyme-substrate complexes. The rate of inactivation by 6-Cl-PRT, which modifies the active site sulfhydryl of cysteine-331, increases with pH; the pK of 7.5 reflects the ionization of the sulfhydryl in the E.6-Cl-PRT complex. The pKs of the acids observed in the IMPDH reaction likely also reflect ionization of the cysteine-331 sulfhydryl which adds to C-2 of IMP prior to NAD reduction. The kcat and kcat/Km values for hydrolysis of 2-Cl-IMP show a pK value of 9.9 for a basic group, similar to that seen in the overall reaction, but do not exhibit the ionization of an acidic group. Surprisingly, the rates of 2-Cl-IMP hydrolysis and of inactivation by 6-Cl-PRT are not stimulated by K+, in contrast to the >100-fold K+ activation of the IMPDH reaction. Apparently the enigmatic role of K+ lies in the NAD(H)-dependent segment of the IMPDH reaction. To evaluate the importance of hydrogen bonding in substrate binding, several deamino- and deoxy-analogues of IMP were tested as substrates and inhibitors. Only 2'-deoxy-IMP was a substrate; the other compounds tested were competitive inhibitors with Ki values at most 10-fold greater than the KD for IMP, illustrating the greater importance of hydrogen-bonding interactions in the chemistry of the IMPDH reaction than simply in nucleotide binding.  相似文献   

11.
The (Na+ +K+)-activated, Mg2+-dependent ATPase from rabbit kidney outer medulla was prepared in a partially inactivated, soluble form depleted of endogenous phospholipids, using deoxycholate. This preparation was reactivated 10 to 50-fold by sonicated liposomes of phosphatidylserine, but not by non-sonicated phosphatidylserine liposomes or sonicated phosphatidylcholine liposomes. The reconstituted enzyme resembled native membrane preparations of (Na+ +K+)-ATPase in its pH optimum being around 7.0, showing optimal activity at Mg2+:ATP mol ratios of approximately 1 and a Km value for ATP of 0.4 mM. Arrhenius plots of this reactivated activity at a constant pH of 7.0 and an Mg2+: ATP mol ratio of 1:1 showed a discontinuity (sharp change of slope) at 17 degrees C, with activation energy (Ea) values of 13-15 kcal/mol above this temperature and 30-35 kcal below it. A further discontinuity was also found at 8.0 degrees C and the Ea below this was very high (greater than 100 kcal/mol). Increased Mg2+ concentrations at Mg2+:ATP ratios in excess of 1:1 inhibited the (Na+ +K+)-ATPase activity and also abolished the discontinuities in the Arrhenius plots. The addition of cholesterol to phosphatidylserine at a 1:1 mol ratio partially inhibited (Na+ +K+)-ATPase reactivation. Arrhenius plots under these conditions showed a single discontinuity at 20 degrees C and Ea values of 22 and 68 kcal/mol above and below this temperature respectively. The ouabain-insensitive Mg2+-ATPase normally showed a linear Arrhenius plot with an Ea of 8 kcal/mol. The cholesterol-phosphatidylserine mixed liposomes stimulated the Mg2+-ATPase activity, which now also showed a discontinuity at 20 degrees C with, however, an increased value of 14 kcal/mol above this temperature and 6 kcal/mol below. Kinetic studies showed that cholesterol had no significant effect on the Km values for ATP. Since both cholesterol and Mg2+ are known to alter the effects of temperature on the fluidity of phospholipids, the above results are discussed in this context.  相似文献   

12.
para-Nitrophenyl phosphorothioate (pNPT) was hydrolyzed by calcineurin at initial rates slightly, but comparable to rates for para-nitrophenyl phosphate (pNPP). Kinetic characterization yielded higher estimates for both Km and Vmax compared to pNPP. Metal ion activation of phosphorothioate hydrolysis was more promiscuous. Unlike the hydrolysis of with pNPP, Ca2+, Mg2+, and Ba2+ activated calcineurin as well as Mn2+.  相似文献   

13.
Sun L  Cui Z  Li C  Huang S  Zhang B 《Biochemistry》2007,46(12):3714-3723
Previously we have identified a highly active ribozyme (R180, cis ribozyme) that can catalyze dipeptide synthesis using N-biotinylcaproyl-aminoacyl-adenylate anhydride (Bio-aa-5'-AMP) as its substrate. In this work, we re-engineered the cis R180 ribozyme into a 158-nt trans ribozyme (TR158) and designed a new substrate (5'-Phe-linker-20-mer). First, the metal ion requirements were examined and compared between the two ribozymes. Both R180 and TR158 ribozymes were active in Mg2+ and Ca2+ but inert with Zn2+, Cu2+, Mn2+, and Co2+. It is intriguing that both ribozymes were highly active in Li+, Na+, or K+ alone but showed very low activity with NH4+. The two ribozymes showed similar linear concentration dependence on Li+ and K+, while they displayed different dependency behavior on Mg2+. Moreover, by using the trans system, the detailed kinetic studies and pH dependent experiments were performed in either 10 mM Mg2+ or 1.0 M Li+. Analysis of kcat and Km values obtained at different pHs (6.0 to 9.0) indicated that it is the catalytic activity of the ribozyme but not the substrate binding affinity that changes significantly with pH. The slopes of the linear parts of the pH-rate plots were close to 1.0 in both Mg2+- and Li+-mediated reactions, suggesting that one proton transfer is involved in the rate-limiting step of catalysis. Overall, our results suggest that Mg2+ and Li+ function similarly in the ribozyme-catalyzed dipeptide synthesis.  相似文献   

14.
S-Adenosylmethionine decarboxylase (AdoMetDC) is a pyruvoyl cofactor-dependent enzyme that participates in polyamine biosynthesis. AdoMetDC from the Archaea Methanococcus jannaschii is a prototype for a recently discovered class that is not homologous to the eucaryotic enzymes or to a distinct group of microbial enzymes. M. jannaschii AdoMetDC has a Km of 95 microm and the turnover number (kcat) of 0.0075 s(-1) at pH 7.5 and 22 degrees C. The turnover number increased approximately 38-fold at a more physiological temperature of 80 degrees C. AdoMetDC was inactivated by treatment with the imine reductant NaCNBH3 only in the presence of substrate. Mass spectrometry of the inactivated protein showed modification solely of the pyruvoyl-containing subunit, with a mass increase corresponding to reduction of a Schiff base adduct with decarboxylated AdoMet. The presteady state time course of the AdoMetDC reaction revealed a burst of product formation; thus, a step after CO2 formation is rate-limiting in turnover. Comparable D2O kinetic isotope effects of were seen on the first turnover (1.9) and on kcat/Km (1.6); there was not a significant D2O isotope effect on kcat, suggesting that product release is rate-limiting in turnover. The pH dependence of the steady state rate showed participation of acid and basic groups with pK values of 5.3 and 8.2 for kcat and 6.5 and 8.3 for kcat/Km, respectively. The competitive inhibitor methylglyoxal bis(guanylhydrazone) binds at a single site per (alphabeta) heterodimer. UV spectroscopic studies show that methylglyoxal bis(guanylhydrazone) binds as the dication with a 23 microm dissociation constant. Studies with substrate analogs show a high specificity for AdoMet.  相似文献   

15.
Francis K  Gadda G 《Biochemistry》2006,45(46):13889-13898
Among the enzymes that catalyze the oxidative denitrification of nitroalkanes to carbonyl compounds, 2-nitropropane dioxygenase is the only one known to effectively utilize both the neutral and anionic (nitronate) forms of the substrate. A recent study has established that the catalytic pathway is common to both types of substrates, except for the initial removal of a proton from the carbon of the neutral substrates [Francis, K., Russell, B., and Gadda, G. (2005) J. Biol. Chem. 280, 5195-5204]. In the present study, the mechanistic properties of the enzyme have been investigated with solvent viscosity, pH, and kinetic isotope effects. With nitroethane or ethylnitronate, the kcat/Km and kcat values were independent of solvent viscosity, consistent with the substrate and product binding to the enzyme in rapid equilibrium. The abstraction of the proton from the alpha carbon of neutral substrates was investigated by measuring the pH dependence of the D(kcat/KNE) value with 1,1-[2H2]-nitroethane. The formation of the enzyme-bound flavosemiquinone formed during catalysis was examined by determining the pH dependence of the kcat/Km values with ethylnitronate and nitroethane and the inhibition by m-nitrobenzoate. Finally, alpha-secondary kinetic isotope effects with 1-[2H]-ethylnitronate were used to propose a non-oxidative tautomerization pathway, in which the enzyme catalyzes the interconversion of nitroalkanes between their anionic and neutral forms. The data presented suggest that enzymatic turnover of 2-nitropropane dioxygenase with neutral substrates is limited by the cleavage of the substrate CH bond at low pH, whereas that with anionic substrates is limited by the non-oxidative tautomerization of ethylnitroante to nitroethane at high pH.  相似文献   

16.
Isotope effects have been measured with Mg2+ as the activator and L-malate labeled with deuterium or tritium at carbon 2 as the substrate over the pH range 4-10. Comparison of the nearly pH-independent deuterium-isotope effect on V/Kmalate of 1.5 with the tritium effect of 2.0 by the method of Northrop (Northrop, D.B. (1975), Biochemistry 14, 2644) gives limits on the true effect of deuterium substitution on the bond-breaking step of 5-8 in the forward reaction and 4-6.5 in the reverse direction. Comparison of the deuterium effect on V/K with the 13C-isotope effect of 1.031 reported by Schimerlik et al. (Schimerlik, M.I., Rife, J.E., and Cleland, W.W. (1975), Biochemistry 14, 5347) allows the deduction that at pH 8 reverse hydride transfer is six to eight times faster than decarboxylation, which is thus largely rate limiting for the catalytic reaction. The absence of a deuterium-isotope effect on V at pH 7-8 and comparison of the Ki of pyruvate as an uncompetitive inhibitor of the forward reaction and a substrate for the reverse reaction indicate that at neutral pH the release of TPNH from enzyme-reduced triphosphopyridine nucleotide (E-TPNH) is the rate-limiting step in the forward direction. The observation of a deuterium effect on V that approaches 3 at pH 4 and 10 shows, however, that, at very low and very high pH, hydride transfer may become partly rate limiting. In the reverse reaction, the probable rate-limiting step at pH 7 is the isomerization of E-TPNH, while at pH 8.5 and above V becomes too large to measure and appears infinite. Substitution of Co2+, Ni2+, or low levels of Mn2+ for Mg2+ gives similar deuterium-isotope effects, although the values of V and Kmalate vary considerably with metal. The kinetics of Mn2+ show pronounced negative cooperativity, with Km values of 7 muM and 5 mM for concentration ranges from 4 to 100 muM and over 1 mM.  相似文献   

17.
1. Transglutaminase (EC 2.3.2.13) was purified from rat liver. 2. The enzyme was stable at 25 degrees C in the pH range of 6.0-9.0, with the optimum at pH 9.0. 3. The enzyme was inactivated after incubation for 20, 4 and 1 min at 44 degrees C, 52 degrees C, and 60 degrees C, respectively. 4. Activation energies were 30.4 kcal/mol for denaturation and 19.9 kcal/mol for substrate conversion to products. 5. The enzyme was inactivated by sulfhydryl modification with hydroxymercuribenzoate (99.1%) and N-ethylmalemide (78.5%). 6. Calcium, required for the activity, was replaced to a lesser extent, by Mg2+, Sr2+, Zn2+ and Mn2+ (31.8, 27.0, 24.6 and 3.5%). 7. Steady-state kinetics showed: Vmax = 10 microM-min-1, Km = 0.05 mM (N-dimethylated casein), kcat = 31.9 min-1 kcat/Km = 560 min-1 mM-1.  相似文献   

18.
Effects of various cations on the dephosphorylation of (Na+ + K+)-ATPase, phosphorylated by ATP in 50 mM imidazole buffer (pH 7.0) at 22 degrees C without added Na+, have been studied. The dephosphorylation in imidazole buffer without added K+ is extremely sensitive to K+-activation (Km K+ = 1 microM), less sensitive to Mg2+-activation (Km Mg2+ = 0.1 mM) and Na+-activation (Km Na+ = 63 mM). Imidazole and Na+ effectively inhibit K+-activated dephosphorylation in linear competitive fashion (Ki imidazole 7.5 mM, Ki Na+ 4.6 mM). The Ki for Na+ is independent of the imidazole concentration, indicating different and non-interacting inhibitory sites for Na+ and imidazole. Imidazole inhibits Mg2+-activated dephosphorylation just as effective as K+-activated dephosphorylation, as judged from the Ki values for imidazole in the two processes. Tris buffer and choline chloride, like imidazole, inhibit dephosphorylation in the presence of residual K+ (less than 1 microM), but less effectively in terms of I50 values and extent of inhibition. Tris inhibits to the same extent as choline. This indicates different inhibitory sites for Tris or choline and for imidazole. These findings indicate that high steady-state phosphorylation levels in Na+-free imidazole buffer are due to the induction of a phosphorylating enzyme conformation and to the inhibition of (K+ + Mg2+)-stimulated dephosphorylation.  相似文献   

19.
The second-order rate constants (kcat/Km) for the beta-glucosidase-catalyzed hydrolysis of aryl beta-D-glucopyranosides show a bell-shaped dependence of pH. The pKas that characterize this dependence are 4.4 (delta Hion approximately equal to 0) and 6.7 (delta Hion approximately equal to 0). In D2O these pKas are increased by 0.5 (+/- 0.1) unit, but there is no solvent isotope effect on the pH-independent second-order rate constant. Nath and Rydon [Nath, R. L., & Rydon, H. N. (1954) Biochem. J. 57, 1-10] examined the kinetics of the beta-glucosidase-catalyzed hydrolysis of a series of substituted phenyl glucosides. We have extended this study to include glucosides with phenol leaving groups of pKa less than 7. Br?nsted plots for this extended series were nonlinear for both kcat/Km and kcat. Br?nsted coefficients for those compounds with leaving groups of pKa greater than 7 (for kcat/Km) or pKa greater than 8.5 (for kcat) were nearly equal to -1.0, indicating substantial negative charge buildup on the leaving group in the transition state. The nonlinearity indicates an intermediate in the reaction. This was confirmed by partitioning experiments in the presence of methanol as a competing glucose acceptor. A constant product ratio, [methyl glucoside]/[glucose], was found with aryl glucoside substrates varying over 16,000-fold in reactivity (V/K), indicative of a common intermediate. Viscosity variation (in sucrose-containing buffers) was used to probe the extent to which the beta-glucosidase reactions are diffusion-controlled.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Coupled spectrophotometric assays that monitor the formation of fumarate and ammonia in the direction of aspartate deamination and aspartate in the direction of fumarate amination were used to collect initial velocity data for the aspartase reaction. Data are consistent with rapid equilibrium ordered addition of Mg2+ prior to aspartate but completely random release of Mg2+, NH4+, or fumarate. In addition to Mg2+, Mn2+ can also be used as a divalent metal with Vmax 80% and a Kaspartate 3.5-fold lower than when Mg2+ is used. Monovalent cations such as Li+, K+, Cs+, and Rb+ are competitive vs. either aspartate or NH4+ but noncompetitive vs. fumarate. A primary deuterium isotope effect of about 1 on both V and V/Kaspartate is obtained with (3R)-L-aspartate-3-d, while a primary 15N isotope effect on V/Kaspartate of 1.0239 +/- 0.0014 is obtained in the direction of aspartate deamination. A secondary isotope effect on V of 1.13 +/- 0.04 is obtained with L-aspartate-2-d. In addition, a secondary isotope effect of 0.81 +/- 0.05 on V is obtained with fumarate-d2, while a value of 1.18 +/- 0.05 on V is obtained by using (2S,3S)-L-aspartate-2,3-d2. These data are interpreted in terms of a two-step mechanism with an intermediate carbanion in which C-N bond cleavage limits the overall rate and the rate-limiting transition state is intermediate between the carbanion and fumarate.  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