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.
Cysteine proteinases are relevant to several aspects of the parasite life cycle and of parasite-host relationships. Here, a quantitative investigation of the effect of temperature and pH on the total substrate inhibition of cruzipain, the major papain-like cysteine proteinase from Trypanosoma cruzi, is reported. Values of the apparent catalytic and inhibition parameters Km, Vmax, Vmax/Km, and K(i) for the cruzipain-catalysed hydrolysis of N-alpha-benzyloxycarbonyl-L-phenylalanyl-L-arginine-(7-amino-4-methylcoumarin) (Z-Phe-Arg-AMC) and azocasein were determined between 10.0 degrees C and 40.0 degrees C and between pH 4.5 and 8.5. Values of Km were independent of temperature and pH, whereas values of Vmax, Vmax/Km, and K(i) were temperature-dependent and pH-dependent. Over the whole pH range explored, values of logVmax, log(Vmax/Km), and logK(i) increased linearly with respect to T(-1). Values of Vmax and Vmax/Km were affected by the acid-base equilibrium of one temperature-independent ionizing group (i.e. pK(unl)' = pK(lig)' = 5.7 +/- 0.1, at 25.0 degrees C). Moreover, values of K(i) were affected by the alkaline pK shift of one ionizing group of active cruzipain (from pK(unl)" = 5.7 +/- 0.1 to pK(lig)" = 6.1 +/- 0.1, at 25.0 degrees C) upon Z-Phe-Arg-AMC binding. Values of logK(unl)', logK(lig)', and logK(lig)" were temperature-independent. Conversely, values of logK(unl)" were linearly dependent on T(-1). As a whole, total substrate inhibition of cruzipain decreased with increasing temperature and pH. These data suggest that both synthetic and protein substrates can bind to the unique active centre of cruzipain either productively or following a binding mode which results in enzyme inhibition. However, allosteric effect(s) cannot be excluded.  相似文献   

2.
M Zolkiewski  A Ginsburg 《Biochemistry》1992,31(48):11991-12000
Dodecameric glutamine synthetase (GS) from Escherichia coli undergoes reversible, thermally induced partial unfolding without subunit dissociation. A single endotherm for Mn.GS (+/- active-site ligands) in the presence of 1 mM free Mn2+ and 100 mM KCl at pH 7 is observed by differential scanning calorimetry (DSC). Previous deconvolutions of DSC data for Mn.GS showed only two two-state transitions (with similar tm values; 51.6 +/- 2 degrees C), and indicated that cooperative interactions link partial unfolding reactions of all subunits within the Mn.enzyme dodecamer [Ginsburg, A., & Zolkiewski, M. (1991) Biochemistry 30, 9421]. A net uptake of 8.0 equiv of H+ by Mn.GS occurs during partial unfolding, as determined in the present DSC experiments conducted with four buffers having different heats of protonation at 50 degrees C. These data gave a value of 176 +/- 12 kcal (mol of dodecamer)-1 for delta Hcal corrected for buffer protonation. L-Glutamine and L-Met-(SR)-sulfoximine stabilize the Mn.GS dodecamer through the free energies of ligand binding, and these were shown to be partially and totally released, respectively, from the 12 active sites at high temperature. Ligand effects on Tm values from DSC were similar to those from spectral measurements of Trp and Tyr exposures in two subunit domains. Effects of varying [ADP] on DSC profiles of Mn.GS were complex; Tm is increased by low [ADP] and decreased by > 100 microM free ADP. This is due to the exposure of an additional low-affinity ADP binding site per GS subunit at high temperature with log K1' = 4.3 and log K2' = 3.6 at 60 degrees C relative to log K' = 5.5 for ADP binding at 30 degrees C, as determined by isothermal calorimetric and fluorescence titrations. Moreover, delta Hcal at > 27% saturation with ADP (corrected for ADP binding/dissociation) is approximately 80-100 kcal/mol more than in the absence of ligands. Changes in domain interactions could result from ADP bridging subunit contacts in the dodecamer. Each of the active-site ligands investigated here produces different effects on DSC profiles without uncoupling the extremely cooperative, partial unfolding reactions in the Mn.GS dodecamer.  相似文献   

3.
The catalytic mechanism underlying the aminopeptidase from Streptomyces griseus (SGAP) was investigated. pH-dependent activity profiles revealed the enthalpy of ionization for the hydrolysis of leucine-para-nitroanilide by SGAP. The value obtained (30 +/- 5 kJ.mol(-1)) is typical of a zinc-bound water molecule, suggesting that the zinc-bound water/hydroxide molecule acts as the reaction nucleophile. Fluoride was found to act as a pure noncompetitive inhibitor of SGAP at pH values of 5.9-8 with a K(i) of 11.4 mM at pH 8.0, indicating that the fluoride ion interacts equally with the free enzyme as with the enzyme-substrate complex. pH-dependent pK(i) experiments resulted in a pK(a) value of 7.0, suggesting a single deprotonation step of the catalytic water molecule to an hydroxide ion. The number of proton transfers during the catalytic pathway was determined by monitoring the solvent isotope effect on SGAP and its general acid-base mutant SGAP(E131D) at different pHs. The results indicate that a single proton transfer is involved in catalysis at pH 8.0, whereas two proton transfers are implicated at pH 6.5. The role of Glu131 in binding and catalysis was assessed by determining the catalytic constants (K(m), k(cat)) over a temperature range of 293-329 degrees K for both SGAP and the E131D mutant. For the binding step, the measured and calculated thermodynamic parameters for the reaction (free energy, enthalpy and entropy) for both SGAP and the E131D mutant were similar. By contrast, the E131D point mutation resulted in a four orders of magnitude decrease in k(cat), corresponding to an increase of 9 kJ.mol(-1) in the activation energy for the E131D mutant, emphasizing the crucial role of Glu131 in catalysis.  相似文献   

4.
Chemical shifts of base and H1' protons of the single-stranded hexamers d(ATTACC) and d(GGTAAT), of the 1:1 mixtures of these complementary hexamers, and of the self-complementary dodecamer d(ATTACCGGTAAT) were measured at various temperatures in aqueous solution. Four different sample concentrations were used in the case of the dodecamer and of the mixture of the complementary hexamers; the individual hexamers were measured at two different DNA concentrations. Absorbance temperature profiles at five different NaCl concentrations were measured for the dodecamer in order to quantify the effect of the ionic strength on the duplex formation. Under suitable conditions of nucleotide concentration, temperature, and ionic strength, the dodecamer adopts either a B-DNA duplex or a hairpin-loop structure. Chemical shift vs temperature profiles, constructed for all samples, were used to obtain thermodynamic parameters either for the various stacking interactions in the single strands or for the duplex or the hairpin-loop formation. In the analysis of the duplex formation of the hexamers, a two-state approach appeared too simple, because systematic deviations were revealed. Therefore, a new three-state model (DUPSTAK) was developed. In order to investigate the magnitude of error arising from the use of the two-state approach in cases where the DUPSTAK model appears more appropriate, a series of test calculations was made. The magnitude of error in the enthalpy and in the entropy of duplex melting is found to depend linearly upon the actual melting temperature and not upon the individual delta Hd degrees and delta Sd degrees values. Thermodynamic analysis of the chemical shift vs temperature profiles in D2O solution (no added salt) yields an average Tmd value of 341 K (1M DNA) and delta Hd degrees of - 121 kJ.mol-1 for the dimer/random-coil transition of the hexamer duplex d(ATTACC).d(GGTAAT). For the duplex in equilibrium random-coil transition of the 12-mer d(ATTACCGGTAAT) an average Tmd value of 336 K (1M DNA) and delta Hd degrees of -372 kJ.mol-1 are found. The hairpin/random-coil transition of d(ATTACCGGTAAT) is characterized by a rather large delta Hh degrees value, -130 kJ.mol-1, and an average Tmh value of 304 K.  相似文献   

5.
To elucidate a detailed catalytic mechanism for nitrile hydratases (NHases), the pH and temperature dependence of the kinetic constants k(cat) and K(m) for the cobalt-type NHase from Pseudonocardia thermophila JCM 3095 (PtNHase) were examined. PtNHase was found to exhibit a bell-shaped curve for plots of relative activity versus pH at pH 3.2-11 and was found to display maximal activity between pH 7.2 and 7.8. Fits of these data provided pK(E)(S1) and pK(E)(S2) values of 5.9 +/- 0.1 and 9.2 +/- 0.1 (k(cat)' = 130 +/- 1 s(-1)), respectively, and pK(E)(1) and pK(E)(2) values of 5.8 +/- 0.1 and 9.1 +/- 0.1 (k(cat)'/K(m)' = (6.5 +/- 0.1) x 10(3) s(-1) mm(-1)), respectively. Proton inventory studies indicated that two protons are transferred in the rate-limiting step of the reaction at pH 7.6. Because PtNHase is stable at 60 degrees C, an Arrhenius plot was constructed by plotting ln(k(cat)) versus 1/T, providing E(a) = 23.0 +/- 1.2 kJ/mol. The thermal stability of PtNHase also allowed DeltaH(0) ionization values to be determined, thus helping to identify the ionizing groups exhibiting the pK(E)(S1) and pK(E)(S2) values. Based on DeltaH(0)(ion) data, pK(E)(S1) is assigned to betaTyr(68), whereas pK(E)(S2) is assigned to betaArg(52), betaArg(157), or alphaSer(112) (NHases are alpha(2)beta(2)-heterotetramers). A combination of these data with those previously reported for NHases and synthetic model complexes, along with sequence comparisons of both iron- and cobalt-type NHases, allowed a novel catalytic mechanism for NHases to be proposed.  相似文献   

6.
The strong ion approach provides a quantitative physicochemical method for describing the mechanism for an acid-base disturbance. The approach requires species-specific values for the total concentration of plasma nonvolatile buffers (A(tot)) and the effective dissociation constant for plasma nonvolatile buffers (K(a)), but these values have not been determined for human plasma. Accordingly, the purpose of this study was to calculate accurate A(tot) and K(a) values using data obtained from in vitro strong ion titration and CO(2) tonometry. The calculated values for A(tot) (24.1 mmol/l) and K(a) (1.05 x 10(-7)) were significantly (P < 0.05) different from the experimentally determined values for horse plasma and differed from the empirically assumed values for human plasma (A(tot) = 19.0 meq/l and K(a) = 3.0 x 10(-7)). The derivatives of pH with respect to the three independent variables [strong ion difference (SID), PCO(2), and A(tot)] of the strong ion approach were calculated as follows: dpH/dSID(+) = [1 + 10(pK(a)-pH)](2)/(2.303 x [SPCO(2)10(pH-pK'(1)[1 + 10(pK(a)-pH](2) + A(tot)10(pK(a)-PH]]; dpH/dPCO(2) = S10(-pK'(1)/[2.303[A(tot)10(pH)(10(pH + 10(pK(a))(-2) - SID(+)10(-pH)]], dpH/dA(tot) = -1/[2.303[SPCO(2)10(pH-pK'(1) + SID(+)10(pK(a)-pH)]], where S is solubility of CO(2) in plasma. The derivatives provide a useful method for calculating the effect of independent changes in SID(+), PCO(2), and A(tot) on plasma pH. The calculated values for A(tot) and K(a) should facilitate application of the strong ion approach to acid-base disturbances in humans.  相似文献   

7.
Co-ordination of catalytic Zn2+ in sorbitol/xylitol dehydrogenases of the medium-chain dehydrogenase/reductase superfamily involves direct or water-mediated interactions from a glutamic acid residue, which substitutes a homologous cysteine ligand in alcohol dehydrogenases of the yeast and liver type. Glu154 of xylitol dehydrogenase from the yeast Galactocandida mastotermitis (termed GmXDH) was mutated to a cysteine residue (E154C) to revert this replacement. In spite of their variable Zn2+ content (0.10-0.40 atom/subunit), purified preparations of E154C exhibited a constant catalytic Zn2+ centre activity (kcat) of 1.19+/-0.03 s(-1) and did not require exogenous Zn2+ for activity or stability. E154C retained 0.019+/-0.003% and 0.74+/-0.03% of wild-type catalytic efficiency (kcat/K(sorbitol)=7800+/-700 M(-1) x s(-1)) and kcat (=161+/-4 s(-1)) for NAD+-dependent oxidation of sorbitol at 25 degrees C respectively. The pH profile of kcat/K(sorbitol) for E154C decreased below an apparent pK of 9.1+/-0.3, reflecting a shift in pK by about +1.7-1.9 pH units compared with the corresponding pH profiles for GmXDH and sheep liver sorbitol dehydrogenase (termed slSDH). The difference in pK for profiles determined in 1H2O and 2H2O solvent was similar and unusually small for all three enzymes (approximately +0.2 log units), suggesting that the observed pK in the binary enzyme-NAD+ complexes could be due to Zn2+-bound water. Under conditions eliminating their different pH-dependences, wild-type and mutant GmXDH displayed similar primary and solvent deuterium kinetic isotope effects of 1.7+/-0.2 (E154C, 1.7+/-0.1) and 1.9+/-0.3 (E154C, 2.4+/-0.2) on kcat/K(sorbitol) respectively. Transient kinetic studies of NAD+ reduction and proton release during sorbitol oxidation by slSDH at pH 8.2 show that two protons are lost with a rate constant of 687+/-12 s(-1) in the pre-steady state, which features a turnover of 0.9+/-0.1 enzyme equivalents as NADH was produced with a rate constant of 409+/-3 s(-1). The results support an auxiliary participation of Glu154 in catalysis, and possible mechanisms of proton transfer in sorbitol/xylitol dehydrogenases are discussed.  相似文献   

8.
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   

9.
l-beta-lysine and beta-glutamate are produced by the actions of lysine 2,3-aminomutase and glutamate 2,3-aminomutase, respectively. The pK(a) values have been titrimetrically measured and are for l-beta-lysine: pK(1)=3.25 (carboxyl), pK(2)=9.30 (beta-aminium), and pK(3)=10.5 (epsilon-aminium). For beta-glutamate the values are pK(1)=3.13 (carboxyl), pK(2)=3.73 (carboxyl), and pK(3)=10.1 (beta-aminium). The equilibrium constants for reactions of 2,3-aminomutases favor the beta-isomers. The pH and temperature dependencies of K(eq) have been measured for the reaction of lysine 2,3-aminomutase to determine the basis for preferential formation of beta-lysine. The value of K(eq) (8.5 at 37 degrees C) is independent of pH between pH 6 and pH 11; ruling out differences in pK-values as the basis for the equilibrium constant. The K(eq)-value is temperature-dependent and ranges from 10.9 at 4 degrees C to 6.8 at 65 degrees C. The linear van't Hoff plot shows the reaction to be enthalpy-driven, with DeltaH degrees =-1.4 kcal mol(-1) and DeltaS degrees =-0.25 cal deg(-1) mol(-1). Exothermicity is attributed to the greater strength of the bond C(beta)-N(beta) in l-beta-lysine than C(alpha)-N(alpha) in l-lysine, and this should hold for other amino acids.  相似文献   

10.
Catalytic mechanism of hamster arylamine N-acetyltransferase 2   总被引:1,自引:0,他引:1  
Wang H  Liu L  Hanna PE  Wagner CR 《Biochemistry》2005,44(33):11295-11306
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from AcCoA to primary arylamines, hydrazines, and hydrazides and play a very important role in the metabolism and bioactivation of drugs, carcinogens, and other xenobiotics. The reaction follows a ping-pong bi-bi mechanism. Structure analysis of bacterial NATs revealed a Cys-His-Asp catalytic triad that is strictly conserved in all known NATs. Previously, we have demonstrated by kinetic and isotope effect studies that acetylation of the hamster NAT2 is dependent on a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)) and not a general acid-base catalysis. In addition, we established that, after formation of the acetylated enzyme intermediate, the active-site imidazole, His-107, is likely deprotonated at physiological pH. In this paper, we report steady-state kinetic studies of NAT2 with two acetyl donors, acetyl coenzyme A (AcCoA) and p-nitrophenyl acetate (PNPA), and four arylamine substrates. The pH dependence of k(cat)/K(AcCoA) exhibited two inflection points at 5.32 +/- 0.13 and 8.48 +/- 0.24, respectively. The pK(a) at 5.32 is virtually identical with the previously reported pK(a) of 5.2 for enzyme acetylation, reaffirming that the first half of the reaction is catalyzed by a thiolate-imidazolium ion pair in the active site. The inflection point at 8.48 indicates that a pH-sensitive group on NAT2 is involved in AcCoA binding. A Br?nsted plot constructed by the correlation of log k(4) and log k(H)2(O) with the pK(a) for each arylamine substrate and water displays a linear free-energy relationship in the pK(a) range from -1.7 (H(2)O) to 4.67 (PABA), with a slope of beta(nuc) = 0.80 +/- 0.1. However, a further increase of the pK(a) from 4.67 (PABA) to 5.32 (anisidine) resulted in a 2.5-fold decrease in the k(4) value. Analysis of the pH-k(cat)/K(PABA) profile revealed a pK(a) of 5.52 +/- 0.14 and a solvent kinetic isotope effect (SKIE) of 2.01 +/- 0.04 on k(cat)/K(PABA). Normal solvent isotope effects of 4.8 +/- 0.1, 3.1 +/- 0.1, and 3.2 +/- 0.1 on the k(cat)/K(b) for anisidine, pABglu, and PNA, respectively, were also determined. These observations are consistent with a deacetylation mechanism dominated by nucleophilic attack of the thiol ester for arylamines with pK(a) values or=5.5. The general base is likely His-107 because the His-107 to Gln and Asn mutants were found to be devoid of catalytic activity. In contrast, an increase in pH-dependent hydrolysis of the acetylated enzyme was not observed over a pH range of 5.2-7.5. On the basis of these observations, a catalytic mechanism for the acetylation of arylamines by NAT2 is proposed.  相似文献   

11.
The unusually low pK(a) value of the general base catalyst Pro-1 (pK(a) = 6.4) in 4-oxalocrotonate tautomerase (4-OT) has been ascribed to both a low dielectric constant at the active site and the proximity of the cationic residues Arg-11 and Arg-39 [Stivers, J. T., Abeygunawardana, C., Mildvan, A. S., Hajipour, G., and Whitman, C. P. (1996) Biochemistry 35, 814-823]. In addition, the pH-rate profiles in that study showed an unidentified protonated group essential for catalysis with a pK(a) of 9.0. To address these issues, the pK(a) values of the active site Pro-1 and lower limit pK(a) values of arginine residues were determined by direct (15)N NMR pH titrations. The pK(a) values of Pro-1 and of the essential acid group were determined independently from pH-rate profiles of the kinetic parameters of 4-OT in arginine mutants of 4-OT and compared with those of wild type. The chemical shifts of all of the Arg Nepsilon resonances in wild-type 4-OT and in the R11A and R39Q mutants were found to be independent of pH over the range 4.9-9.7, indicating that no arginine is responsible for the kinetically determined pK(a) of 9.0 for an acidic group in free 4-OT. With the R11A mutant, where k(cat)/K(m) was reduced by a factor of 10(2.9), the pK(a) of Pro-1 was not significantly altered from that of the wild-type enzyme (pK(a) = 6.4 +/- 0.2) as revealed by both direct (15)N NMR titration (pK(a) = 6.3 +/- 0.1) and the pH dependence of k(cat)/K(m) (pK(a) = 6.4 +/- 0.2). The pH-rate profiles of both k(cat)/K(m) and k(cat) for the reaction of the R11A mutant with the dicarboxylate substrate, 2-hydroxymuconate, showed humps, i.e., sharply defined maxima followed by nonzero plateaus. The humps disappeared in the reaction with the monocarboxylate substrate, 2-hydroxy-2,4-pentadienoate, indicating that, unlike the wild-type enzyme which reacts only with the dianionic form of the dicarboxylic substrate, the R11A mutant reacts with both the 6-COOH and 6-COO(-) forms, with the 6-COOH form being 12-fold more active. This reversal in the preferred ionization state of the 6-carboxyl group of the substrate that occurs upon mutation of Arg-11 to Ala provides strong evidence that Arg-11 interacts with the 6-carboxylate of the substrate. In the R39Q mutant, where k(cat)/K(m) was reduced by a factor of 10(3), the kinetically determined pK(a) value for Pro-1 was 4.6 +/- 0.2, while the ionization of Pro-1 showed negative cooperativity with an apparent pK(a) of 7.1 +/- 0.1 determined by 1D (15)N NMR. From the Hill coefficient of 0.54, it can be shown that the apparent pK(a) value of 7.1 could result most simply from the averaging of two limiting pK(a) values of 4.6 and 8.2. Mutation of Arg-39, by altering the structure of the beta-hairpin which covers the active site, could result in an increase in the solvent exposure of Pro-1, raising its upper limit pK(a) value to 8.2. In the R39A mutant, the kinetically determined pK(a) of Pro-1 was also low, 5.0 +/- 0.2, indicating that in both the R39Q and R39A mutants, only the sites with low pK(a) values were kinetically operative. With the fully active R61A mutant, the kinetically determined pK(a) of Pro-1 (pK(a) = 6.5 +/- 0.2) agreed with that of wild-type 4-OT. It is concluded that the unusually low pK(a) of Pro-1 shows little contribution from electrostatic effects of the nearby cationic Arg-11, Arg-39, and Arg-61 residues but results primarily from a site of low local dielectric constant.  相似文献   

12.
The quantitative mechanistic acid-base approach to clinical assessment of acid-base status requires species-specific values for [A]tot (the total concentration of nonvolatile buffers in plasma) and Ka (the effective dissociation constant for weak acids in plasma). The aim of this study was to determine [A]tot and Ka values for plasma in domestic pigeons. Plasma from 12 healthy commercial domestic pigeons was tonometered with 20% CO2 at 37 degrees C. Plasma pH, Pco2, and plasma concentrations of strong cations (Na, K, Ca), strong anions (Cl, L-lactate), and nonvolatile buffer ions (total protein, albumin, phosphate) were measured over a pH range of 6.8-7.7. Strong ion difference (SID) (SID5=Na+K+Ca-Cl-lactate) was used to calculate [A]tot and Ka from the measured pH and Pco2 and SID5. Mean (+/-SD) values for bird plasma were as follows: [A]tot=7.76+/-2.15 mmol/l (equivalent to 0.32 mmol/g of total protein, 0.51 mmol/g of albumin, 0.23 mmol/g of total solids); Ka=2.15+/-1.15x10(-7); and pKa=6.67. The net protein charge at normal pH (7.43) was estimated to be 6 meq/l; this value indicates that pigeon plasma has a much lower anion gap value than mammals after adjusting for high mean L-lactate concentrations induced by restraint during blood sampling. This finding indicates that plasma proteins in pigeons have a much lower net anion charge than mammalian plasma protein. An incidental finding was that total protein concentration measured by a multianalyzer system was consistently lower than the value for total solids measured by refractometer.  相似文献   

13.
D M Quinn 《Biochemistry》1985,24(13):3144-3149
Solvent deuterium isotope effects on the rates of lipoprotein lipase (LpL) catalyzed hydrolysis of the water-soluble esters p-nitrophenyl acetate (PNPA) and p-nitrophenyl butyrate (PNPB) have been measured and fall in the range 1.5-2.2. The isotope effects are independent of substrate concentration, LpL stability, and reaction temperature and hence are effects on chemical catalysis and not due to a medium effect of D2O on LpL stability and/or conformation. pL (L = H or D) vs. rate profiles for the Vmax/Km of LpL-catalyzed hydrolysis of PNPB increase sigmoidally with increasing pL. Least-squares analysis of the profiles gives pKaH2O = 7.10 +/- 0.01, pKaD2O = 7.795 +/- 0.007, and a solvent isotope effect on limiting velocity at high pL of 1.97 +/- 0.03. Because the pL-rate profiles are for the Vmax/Km of hydrolysis of a water-soluble substrate, the measured pKa's are intrinsic acid-base ionization constants for a catalytically involved LpL active-site amino acid side chain. Benzeneboronic acid, a potent inhibitor of LpL-catalyzed hydrolysis of triacylglycerols [Vainio, P., Virtanen, J. A., & Kinnunen, P. K. J. (1982) Biochim. Biophys. Acta 711, 386-390], inhibits LpL-catalyzed hydrolysis of PNPB, with Ki = 6.9 microM at pH 7.36, 25 degrees C. This result and the solvent isotope effects for LpL-catalyzed hydrolysis of water-soluble esters are interpreted in terms of a proton transfer mechanism that is similar in many respects to that of the serine proteases.  相似文献   

14.
Earlier studies have reported that trimethylamine N-oxide (TMAO), a naturally occurring osmolyte, is a universal stabilizer of proteins because it folds unstructured proteins and counteracts the deleterious effects of urea and salts on the structure and function of proteins. This conclusion has been reached from the studies of the effect of TMAO on proteins in the pH range 6.0-8.0. In this pH range TMAO is almost neutral (zwitterionic form), for it has a pK(a) of 4.66 +/- 0.10. We have asked the question of whether the effect of TMAO on protein stability is pH-dependent. To answer this question we have carried out thermal denaturation studies of lysozyme, ribonuclease-A, and apo-alpha-lactalbumin in the presence of various TMAO concentrations at different pH values above and below the pK(a) of TMAO. The main conclusion of this study is that near room temperature TMAO destabilizes proteins at pH values below its pK(a), whereas it stabilizes proteins at pH values above its pK(a). This conclusion was reached by determining the T(m) (midpoint of denaturation), delta H(m) (denaturational enthalpy change at T(m)), delta C(p) (constant pressure heat capacity change), and delta G(D) degrees (denaturational Gibbs energy change at 25 degrees C) of proteins in the presence of different TMAO concentrations. Other conclusions of this study are that T(m) and delta G(D) degrees depend on TMAO concentration at each pH value and that delta H(m) and the delta C(p) are not significantly changed in presence of TMAO.  相似文献   

15.
The acid-base properties of Adenosine 5'-triphosphate (ATP) in NaCl and KCl aqueous solutions at different ionic strengths (0相似文献   

16.
Equilibrium and calorimetric studies of substrate binding to turkey egg white (TEW) lysozyme were carried out at 30degrees as a function of pH (2 to 9) and ligand size (monosaccharide to hexasaccharide of N-acetylglucosamine). Steady state kinetic measurements using the N-acetylglucosamine hexasaccharide were carried out as a function of pH (2 to 9) and temperature (20-60degrees). These experiments allow comparison of the properties of TEW lysozyme with those of the hen egg white (HEW) enzyme reported previously (Banerjee, S. K., Holler, E., Hess, G. P., and Rupley, J. A. (1975) J. Biol. Chem. 250, 4355-4367, and references therein). The free energies and enthalpies of oligosaccharide binding are the same for TEW and HEW lysozymes at pH 2 but are less negative for TEW lysozyme at pH 5. The pH dependence of the binding of (GlcNAc)3 and higher oligomers to TEW lysozyme is like that for the binding of beta-methyl-N-acetylglucosaminide to TEW lysozyme. These data indicate that oligosaccharide ligands bind identically with HEW and TEW lysozymes, except for the interactions of residue 101, which is aspartic acid in the HEW protein and glycine in the TEW protein (Larue, J. N., and Speck, J. C., Jr. (1970) J. Biol. Chem. 245, 1985-1991). The pH dependence of kcat is described by apparent pK values of 3.9 and 6.8 and a maximum value of kcat of 0.135 s-1. A value of 21.0 kcal/mol was calculated for deltaH from the temperature dependence of kcat. These values and the dependence of the transglycosylation reaction on acceptor concentration are within experimental error the same as those for HEW lysozyme. The more acid pK seen in the pH rate profile reflects the ionization of Asp-52 in the lysozyme-(GlcNAc)6 complex. The pK of Asp-52 in the free protein is 0.3 pK unit lower. The essential identity of the active sites of the HEW and TEW enzymes, except for the Asp-101 interactions, allows estimation of the thermodynamic properties associated with formation of the two hydrogen bonds between Asp-101 and substrate as deltaG0 = -1.2 kcal/mol, DeltaH0 = -3.6 kcal/mol, and deltaS0 = -7.9 e.u.  相似文献   

17.
A model, developed within the framework of the counterion condensation theory of linear polyelectrolytes, is presented in this paper to describe the acid-base properties of linear polyelectrolytes, consisting of several types of functional ionizable groups. This formalism has been successfully applied to Fluka humic acid under salt-free conditions, as well as in the presence of supporting simple 1:1 salt (KNO3) at three different concentrations. As part of this approach, the charge density of the humic acid is obtained from the activity coefficient measurements of potassium counterions at different humic acid concentrations at a constant degree of dissociation of the polyelectrolyte. The humic acid average charge density was found to be 0.80 +/- 0.05. Using the present model, we are able to satisfactorily describe the experimental data obtained from acid-base potentiometric titrations. Four main functional groups making up the polymer are determined through their fractional abundances (Xi) and intrinsic pK (pK0i) values. The fractional abundances remained constant and independent of the ionic strength, indicating that the humic acid constitution does not depend on the concentration of excess salts. The pK0i values show a small change with ionic strength, which can be explained by the polyelectrolytic behavior of the solution.  相似文献   

18.
Marti DN  Jelesarov I  Bosshard HR 《Biochemistry》2000,39(42):12804-12818
Residues of opposite charge often populate heptad positions g (heptad i on chain 1) and e' (heptad i + 1 on chain 2) in dimeric coiled coils and may stabilize the dimer by formation of interchain ion pairs. To investigate the contribution to stability of such electrostatic interactions we have designed a disulfide-linked heterodimeric zipper (AB zipper) consisting of the acidic chain Ac-E-VAQLEKE-VAQAEAE-NYQLEQE-VAQLEHE-CG-NH(2) and the basic chain Ac-E-VQALKKR-VQALKAR-NYAAKQK-VQALRHK-CG-NH(2) in which all e and g positions are occupied by either E or K/R to form a maximum of seven interhelical salt bridges. Temperature-induced denaturation experiments monitored by circular dichroism reveal a stable coiled coil conformation below 50 degrees C and in the pH range 1.2-8.0. Stability is highest at pH approximately 4.0 [DeltaG(U) (37 degrees C) = 5.18 +/- 0.51 kcal mol(-)(1)]. The solution structure of the AB zipper at pH 5.65 has been elucidated on the basis of homonuclear (1)H NMR data collected at 800 MHz [heavy atom rmsd's for the ensemble of 50 calculated structures are 0.47 +/- 0.13 A (backbone) and 0.95 +/- 0.16 A (all)]. Both chains of the AB zipper are almost entirely in alpha-helical conformation and form a superhelix with a left-handed twist. Overhauser connectivities reveal close contacts between g position residues (heptad i on chain 1) and residues d/f (heptad i on chain 1), residues a/d (heptad i + 1 on chain 1), and residue a' (heptad i + 1 on chain 2). Residues in position e (heptad i on chain 1) are in contact with residues a/b/d/f (heptad i on chain 1) and residue d' (heptad i on chain 2). These connectivities hint at a relatively defined alignment of the side chains across the helix interface. Partial H-bond formation between the functional groups of residues g and e'(+1) is observed in the calculated structures. NMR pH titration experiments disclose pK(a) values for Glu delta-carboxylate groups: 4.14 +/- 0.02 (E(1)), 4.82 +/- 0.07 (E(6)), 4.52 +/- 0.01 (E(8)), 4.37 +/- 0.03 (E(13)), 4.11 +/- 0.02 (E(15)), 4.41 +/- 0.07 (E(20)), 4.82 +/- 0.03 (E(22)), 4.65 +/- 0.04 (E(27)), 4.63 +/- 0.03 (E(29)), 4.22 +/- 0.02 (E(1)(')). By comparison with pK(a) of Glu in unfolded peptides ( approximately 4. 3 +/- 0.1), our pK(a) data suggest marginal or even unfavorable contribution of charged Glu to the stability of the AB zipper. The electrostatic energy gained from interhelical ion pairs is likely to be surpassed by hydrophobic energy terms upon protonation of Glu, due to increased hydrophobicity of uncharged Glu and, thus, better packing against apolar residues at the chain interface.  相似文献   

19.
The pH dependence of kcat/Km for the papain-catalyzed hydrolysis of ethyl hippurate, N-alpha-benzoyl-L-citrulline methyl ester, and the p-nitroanilide, amide, and ethyl ester derivatives of N-alpha-benzoyl-L-arginine was determined below pH 6.4. The value of kcat/Km was observed to be modulated by two acid ionizations rather than a single ionization as previously believed. For the five substrates studied, the average pK values for the two ionizations are 3.78 +/- 0.2 and 3.95 +/- 0.1 at T/2 0.3, 25 degrees C. The observation that similar pK values were obtained with different substrates was taken as evidence that the kinetically determined pK values are close in value to true macroscopic ionization constants for ionization of groups on the free enzyme.  相似文献   

20.
The nucleoside analog 2',3'-dideoxycytidine (ddCyd) has been shown to inhibit the infectivity and cytopathic effect of human immunodeficiency virus on human OKT4+ lymphocytes in vitro. Metabolism of ddCyd by human T-lymphoblastic cells (Molt 4) negative for human immunodeficiency virus and OKT4 was examined. Molt 4 cells accumulated ddCyd and its phosphorylated derivatives into acid-soluble and acid-insoluble material in a dose-dependent manner. For each concentration tested, 2',3'-dideoxycytidine triphosphate represented 40% of the total acid-soluble pool of ddCyd metabolites. Uptake of 5 microM ddCyd was linear for 4 h after addition of drug. Efflux of ddCyd metabolites from cells followed a biphasic course with an initial retention half-life of 2.6 h for 2',3'-dideoxycytidine triphosphate. DNA, but not RNA, of cells incubated with [3H]ddCyd became radiolabeled. Nuclease and phosphatase treatment of DNA followed by reverse-phase high pressure liquid chromatography showed that the nucleoside was incorporated into DNA in its original form. ddCyd was not susceptible to deamination by human Cyd-dCyd deaminase. It was a poor substrate for human cytoplasmic and mitochondrial dCyd kinases, with Km values of 180 +/- 30 and 120 +/- 20 microM, respectively. DNA polymerases alpha, beta, and gamma varied in their sensitivity to inhibition by ddCTP with Ki values of 110 +/- 40, 2.6 +/- 0.3, and 0.016 +/- 0.008 microM, respectively; however, inhibition was competitive with dCTP in each case.  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