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 62 毫秒
1.
The effects of salts on the rate constants of inactivation by heat of yeast alcohol dehydrogenase (YADH) at 60.0 degrees C were measured. Different effects were observed at low and high salt concentrations. At high concentrations, some salts had stabilizing effects, while others were destabilizing. The effects of salts in the high concentration range examined can be described as follows: (decreased thermal stability) NaClO(4) < NaI = (C(2)H(5))(4)NBr < NH(4)Br < NaBr = KBr = CsBr = (no addition) < (CH(3))(4)NBr < KCl < KF < Na(2)SO(4) (increased thermal stability). The decreasing effect of NaClO(4) on YADH controlled the thermal stability of the enzyme absolutely and was not compensated by the addition of Na(2)SO(4), a salt which stabilized the enzyme. However, Na(2)SO(4) compensation did occur in response to the decrease in thermal stability caused by (C(2)H(5))(4)NBr. The rate constants of inactivation by heat (k (in)) of the enzyme were measured at various temperatures. Effective values of the thermodynamic activation parameters of thermal inactivation, activation of free energy (DeltaG (double dagger)), activation enthalpy (DeltaH (double dagger)), and activation entropy (DeltaS (double dagger)), were determined. The thermal stability of YADH in 0.8 M Na(2)SO(4) increased more than that of pyruvate kinase from Bacillus stearothermophilus, a moderate thermophile. The changes in the values of DeltaH (double dagger) and DeltaS (double dagger) were great and showed a general compensatory tendency, with the exception of in the case of NaClO(4). The temperature for the general compensation effect (T (c)) was approximately 123 degrees C. With Na(2)SO(4), the thermal stability of YADH at a temperature below T (c) was greater than that in the absence of salt due to the higher values of DeltaH (double dagger) and DeltaS (double dagger), respectively, and thus was an example of low-temperature enzymatic stabilization. With (C(2)H(5))(4)NBr, the thermal stability of YADH at a temperature below T (c) was lower than that in the absence of salt due to the lower values of DeltaH (double dagger) and DeltaS (double dagger), respectively, and thus was an example of low-temperature enzymatic destabilization. But with NaClO(4), the changes in the values of DeltaH (double dagger) and DeltaS (double dagger) were small and the thermal stability of YADH was thus an example of high-temperature enzymatic destabilization.  相似文献   

2.
CFTR is the only ABC (ATP-binding cassette) ATPase known to be an ion channel. Studies of CFTR channel function, feasible with single-molecule resolution, therefore provide a unique glimpse of ABC transporter mechanism. CFTR channel opening and closing (after regulatory-domain phosphorylation) follows an irreversible cycle, driven by ATP binding/hydrolysis at the nucleotide-binding domains (NBD1, NBD2). Recent work suggests that formation of an NBD1/NBD2 dimer drives channel opening, and disruption of the dimer after ATP hydrolysis drives closure, but how NBD events are translated into gate movements is unclear. To elucidate conformational properties of channels on their way to opening or closing, we performed non-equilibrium thermodynamic analysis. Human CFTR channel currents were recorded at temperatures from 15 to 35 degrees C in inside-out patches excised from Xenopus oocytes. Activation enthalpies(DeltaH(double dagger)) were determined from Eyring plots. DeltaH(double dagger) was 117 +/- 6 and 69 +/- 4 kJ/mol, respectively, for opening and closure of partially phosphorylated, and 96 +/- 6 and 73 +/- 5 kJ/mol for opening and closure of highly phosphorylated wild-type (WT) channels. DeltaH(double dagger) for reversal of the channel opening step, estimated from closure of ATP hydrolysis-deficient NBD2 mutant K1250R and K1250A channels, and from unlocking of WT channels locked open with ATP+AMPPNP, was 43 +/- 2, 39 +/- 4, and 37 +/- 6 kJ/mol, respectively. Calculated upper estimates of activation free energies yielded minimum estimates of activation entropies (DeltaS(double dagger)), allowing reconstruction of the thermodynamic profile of gating, which was qualitatively similar for partially and highly phosphorylated CFTR. DeltaS(double dagger) appears large for opening but small for normal closure. The large DeltaH(double dagger) and DeltaS(double dagger) (TDeltaS(double dagger) >/= 41 kJ/mol) for opening suggest that the transition state is a strained channel molecule in which the NBDs have already dimerized, while the pore is still closed. The small DeltaS(double dagger) for normal closure is appropriate for cleavage of a single bond (ATP's beta-gamma phosphate bond), and suggests that this transition state does not require large-scale protein motion and hence precedes rehydration (disruption) of the dimer interface.  相似文献   

3.
Kinetics for the breakdown of the trinuclear chromium acetate cluster, [Cr(3)O(OAc)(6)](+), with a series of monoprotic and diprotic ligands in weakly acidic aqueous media (pH approximately 4 or approximately 5) have been investigated spectrophotometrically at 40-60 degrees C. The results point to an ion-pair equilibrium as the first step followed by associative interchange mechanism forming the mononuclear product of the reaction. Pseudo-first-order rates were determined from absorbance data and associated activation parameters were calculated using the Eyring equation. Enthalpy and entropy terms of the reactions (e.g., histidine, DeltaH(double dagger) = 75 +/- 15 kJ mol(-1), DeltaS(double dagger) = -130 +/- 25 J K(-1) mol(-1); lactic acid, DeltaH(double dagger) = 66 +/- 13 kJ mol(-1), DeltaS(double dagger) = -155 +/- 30 J K(-1) mol(-1); glycine, DeltaH(double dagger) = 31 +/- 6 kJ mol(-1), DeltaS(double dagger) = -225 +/- 45 J K(-1) mol(-1)) are consistent with an associative interchange (I(a)) mechanism, and produce a linear isokinetic plot (slope = 50 degrees C). Rates and activation parameters are comparable to those of substitution reactions of the chromium(III) hexaaqua cation. Other ligands studied included malonic acid and the amino acid, aspartic acid. Observed rates are faster than water exchange rates, but typically slower than anion substitution rates, and indicate that trinuclear chromium(III) clusters are expected to be kinetically stable in neutral to slightly acidic conditions.  相似文献   

4.
The catalytically competent Mn(II)-loaded form of the argE-encoded N-acetyl-l-ornithine deacetylase from Escherichia coli (ArgE) was characterized by kinetic, thermodynamic, and spectroscopic methods. Maximum N-acetyl-l-ornithine (NAO) hydrolytic activity was observed in the presence of one Mn(II) ion with k cat and K m values of 550 s−1 and 0.8 mM, respectively, providing a catalytic efficiency (k cat/K m) of 6.9 × 105 M−1 s−1. The ArgE dissociation constant (K d) for Mn(II) was determined to be 0.18 μM, correlating well with a value obtained by isothermal titration calorimetry of 0.30 μM for the first metal binding event and 5.3 μM for the second. An Arrhenius plot of the NAO hydrolysis for Mn(II)-loaded ArgE was linear from 15 to 55 °C, suggesting the rate-limiting step does not change as a function of temperature over this range. The activation energy, determined from the slope of this plot, was 50.3 kJ mol−1. Other thermodynamic parameters were ΔG = 58.1 kJ mol−1, ΔH = 47.7 kJ mol−1, and ΔS = –34.5 J mol−1 K−1. Similarly, plots of lnK m versus 1/T were linear, suggesting substrate binding is controlled by a single step. The natural product, [(2S,3R)-3-amino-2-hydroxy-4-phenylbutanoyl]leucine (bestatin), was found to be a competitive inhibitor of ArgE with a K i value of 67 μM. Electron paramagnetic resonance (EPR) data recorded for both [Mn(II)_(ArgE)] and [Mn(II)Mn(II)(ArgE)] indicate that the two Mn(II) ions form a dinuclear site. Moreover, the EPR spectrum of [Mn(II)Mn(II)(ArgE)] in the presence of bestatin indicates that bestatin binds to ArgE but does not form a μ-alkoxide bridge between the two metal ions.  相似文献   

5.
Crystalline soybean trypsin inhibitor protein undergoes denaturation on heating which is reversed on cooling. In the range of temperature of 35 to 50 degrees C. a solution of the protein consists of a mixture of native and denatured forms in equilibrium with each other. The equilibrium is only slowly established and its final value at any temperature is the same whether a heated, denatured solution of the protein is cooled to the given temperature or whether a fresh solution is raised to that temperature. The kinetics of reversible denaturation of the soybean protein as well as the reversal of denaturation is that of a reversible unimolecular reaction, each process consisting at a given temperature of the same two simultaneous reactions acting in opposite directions. The experimental data on the effect of temperature on the velocity and the equilibrium constants of the opposing reaction were utilized in evaluating the reaction energies and activation energies. The reaction energies for denaturation were found to be as follows:- Change in total heat of reaction DeltaH = 57,000 calories per mole Change in entropy of reaction DeltaS = 180 calories per degree per mole The heat of activation DeltaH(1) (double dagger) for denaturation = 55,000 The heat of activation DeltaH(2) (double dagger) for the reversal of denaturation = -1900 The entropy DeltaS(1) (double dagger) for denaturation = 95 The entropy DeltaS(2) (double dagger) for reversal of denaturation = -84  相似文献   

6.
Pascher T 《Biochemistry》2001,40(19):5812-5820
Utilizing the stability difference between the ferro and ferri forms of horse heart cytochrome c (cyt c), folding of reduced cyt c was triggered by laser-induced reduction of unfolded oxidized cyt c. Measurements were made of the kinetics of the main folding phase (1 ms-10 s) in which collapsed reduced cyt c transforms to the native conformation. The folding rates were studied extensively as a function of temperature (5-75 degrees C) and guanidine hydrochloride (GdnHCl) concentration (1.6-4.9 M). At constant [GdnHCl], the Arrhenius plot of the folding rate constant (k) is nonlinear. At temperatures above 40 degrees C, the decrease in protein stability counteracts the expected increase in folding rate. Introducing free energy (DeltaG), derived from protein stability data, into the Eyring and Arrhenius equations leads to: ln k = ln(k(b)T/h) + DeltaS()/R - DeltaH()/RT - theta(m)DeltaG/RT = ln A - E(a)/RT - theta(m)DeltaG/RT, where theta(m) is the ratio between the denaturant dependence of the folding rate and the stability. By using this equation at constant DeltaG [or constant equilibrium constant (K)], linear Arrhenius plots are obtained. For the main folding phase of reduced cyt c, a positive DeltaS() is obtained indicating that the transition state is less ordered than the reactant. A model is proposed in which reduced cyt c first collapses into a compact intermediate, which needs to expand to reach the transition state of the rate-limiting folding reaction.  相似文献   

7.
Death of Saccharomyces cerevisiae induced by sulfur dioxide (K(2)S(2)O(2) was used as the SO(2) source) followed saturation kinetics. The enthalpy of activation of death was not affected by concentration over the range tested (5-150) mg/L of (K(2)S(2)O(2) at pH 3.4) and averaged 3.6 x 10(4) cal/mol as compared with 8.5 x 10(4) cal/mol for DeltaH(double dagger) of thermal death. The entropy of activation of death was hyperbolic function of the sulfur dioxide concentration, extrapolated at zero concentration to DeltaS(0) (double dagger) = 36.8 cal mol(-1) K(-1) and tended to DeltaDeltaS(max) (double dagger) = 13.2 cal mol(-1) K(-1) at saturating concentration, yielding a dis0sociation constant of 5.8 x 10(-1) M sulfur dioxide. As was predicted from these results, In K(d) (the specific rate of death induced by sulfur dioxide) was hyperbolic function of concentration under isothermic conditions and extrapolated toa finite value at zero concentration. The Arrhenuis plots and the DeltaS(+/-) plot versus concentration revealed the occurrence of substrate inhabitation of the death effect at high concentrations (above 60 mg/L K(2)S(2)O(2) at pH 3.4). A model is presented involving two types of receptor sites for sulfur dioxide on the cell surface, on directly connected with the death process, the other modulating its entropy of activation.  相似文献   

8.
Described are further studies directed towards elucidating the mechanism of the nitric oxide reduction of the copper(II) model system, Cu(dmp)2(2+) (I, dmp=2,9-dimethyl-1,10-phenanthroline). The reaction of I with NO in methanol results in the formation of Cu(dmp)2+ (II) and methyl nitrite (CH3ONO), with a second order rate constant kNO=38.1 M-1 s-1 (298K). The activation parameters for this reaction in buffered aqueous medium were measured to be DeltaH(double dagger)=41.6 kJ/mol and DeltaS(double dagger)=-82.7 kJ/mol deg. The addition of azide ion (N3-) as a competing nucleophile results in a marked acceleration in the rate of the copper(II) reduction. Analysis of the kinetics for the NO reduction of the bulkier Cu(dpp)(2)2+ (IV, dpp=2,9-diphenyl-1,10-phenanthroline) and the stronger oxidant, Cu(NO2-dmp)2(2+) (V, NO2-dmp=5-nitro-2,9-dimethyl-1,10-phenanthroline), gave the second order rate constants kNO=21.2 and 29.3 M-1 s-1, respectively. These results argue against an outer sphere electron transfer pathway and support a mechanism where the first step involves the formation of a copper-nitrosyl (Cu(II)-NO or Cu(I)-NO+) adduct. This would be followed by the nucleophilic attack on the bound NO and the labilization of RONO to form the nitrite products and the cuprous complex.  相似文献   

9.
Bai TC  Yan GB 《Carbohydrate research》2003,338(24):2921-2927
Viscosity and density data for the system of heptanedioic acid dissolved in aqueous sucrose solution at temperature range from 288.15 to 313.15 K have been measured. The viscosity B-coefficients for heptanedioic acid in aqueous sucrose solution has been calculated. The effect of temperature and sucrose concentration on the B-coefficients is discussed. On the basis of the Feakins equation, the activation parameters (Deltamu3++, DeltaH3++, DeltaS3++, DeltaG12(0)++, DeltaH12(0)++ and DeltaS12(0)++) for viscous flow of the solution have been evaluated, together with the Gibbs energy of transfer for the solute from the ground state solvent to the hypothetical viscous transition state solvent (DeltaG3++(1-1')). The effect of sucrose concentration and temperature on the activation parameters has been discussed.  相似文献   

10.
Interaction kinetic and thermodynamic analyses provide information beyond that obtained in general inhibition studies, and may contribute to the design of improved inhibitors and increased understanding of molecular interactions. Thus, a biosensor-based method was used to characterize the interactions between HIV-1 protease and seven inhibitors, revealing distinguishing kinetic and thermodynamic characteristics for the inhibitors. Lopinavir had fast association and the highest affinity of the tested compounds, and the interaction kinetics were less temperature-dependent as compared with the other inhibitors. Amprenavir, indinavir and ritonavir showed non-linear temperature dependencies of the kinetics. The free energy, enthalpy and entropy (DeltaG, DeltaH, DeltaS) were determined, and the energetics of complex association (DeltaG(on), DeltaH(on), DeltaS(on)) and dissociation (DeltaG(off), DeltaH(off), DeltaS(off)) were resolved. In general, the energetics for the studied inhibitors was in the same range, with the negative free energy change (DeltaG < 0) due primarily to increased entropy (DeltaS > 0). Thus, the driving force of the interaction was increased degrees of freedom in the system (entropy) rather than the formation of bonds between the enzyme and inhibitor (enthalpy). Although the DeltaG(on) and DeltaG(off) were in the same range for all inhibitors, the enthalpy and entropy terms contributed differently to association and dissociation, distinguishing these phases energetically. Dissociation was accompanied by positive enthalpy (DeltaH(off) > 0) and negative entropy (DeltaS(off) < 0) changes, whereas association for all inhibitors except lopinavir had positive entropy changes (DeltaS(on) > 0), demonstrating unique energetic characteristics for lopinavir. This study indicates that this type of data will be useful for the characterization of target-ligand interactions and the development of new inhibitors of HIV-1 protease.  相似文献   

11.
Measurements are reported on the kinetics of the lysozyme-catalysed hydrolysis of several beta-aryl di-N-acetylchitobiosides, some of which have been synthesized for the first time. The catalytic rate constants (k(cat.)) at 45 degrees yield a curved Hammett plot (concave up) and the plot of DeltaH(double dagger) versus DeltaS(double dagger) has a sharp break. Substrates with electron-withdrawing groups exhibit a kinetic deuterium isotope effect (k(H) (cat.)/k(D) (cat.)), whereas those with electron-donating groups show no such isotope effect. The results suggest the operation of different mechanisms for the two types of substrates.  相似文献   

12.
Manganese(II) activation of the methionyl aminopeptidases from Escherichia coli (EcMetAP-I) and the hyperthermophilic archaeon Pyrococcus furiosus (PfMetAP-II) was investigated. Maximum catalytic activity for both enzymes was obtained with 1 equiv of Mn(II), and the dissociation constants (K(d)) for the first metal binding site were found to be 6 +/- 0.5 and 1 +/- 0.5 microM for EcMetAP-I and PfMetAP-II, respectively. These K(d) values were verified by isothermal titration calorimetry (ITC) and found to be 3.0 +/- 0.2 and 1.4 +/- 0.2 microM for EcMetAP-I and PfMetAP-II, respectively. The hydrolysis of MGMM was measured in triplicate between 25 and 85 degrees C at eight substrate concentrations ranging from 2 to 20 mM for PfMetAP-II. Both specific activity and K(m) values increased with increasing temperature. An Arrhenius plot was constructed from the kcat values and was found to be linear over the temperature range 25-85 degrees C. The activation energy for the Mn(II)-loaded PfMetAP-II hydrolysis of MGMM was found to be 25.7 kJ/mol while the remaining thermodynamic parameters calculated at 25 degrees C are DeltaG+ = 50.1 kJ/mol, DeltaH+ = 23.2 kJ/mol, and DeltaS++ = -90.2 J x mol(-1) x K(-1).  相似文献   

13.
The position of the amino acid substitution in the human red cell carbonic anhydrase I variant, CA Id Michigan, has been determined by sequence analysis of the altered tryptic peptide. The threonine to lysine substitution was found to be located at position 100, and is expressed as CA I100 Lys.  相似文献   

14.
Trapp O  Schurig V 《Chirality》2002,14(6):465-470
The axially chiral allenes dimethyl-1,3-allenedicarboxylate 1 and diethyl-1,3-allenedicarboxylate 2 show characteristic plateau formation during enantioselective GC separation on the chiral stationary liquid phase Chirasil-beta-Dex. The elution profiles, obtained from temperature-dependent dynamic GC (DGC) experiments (1: 100-140 degrees C; 2: 110-150 degrees C) were evaluated with the recently derived approximation function (AF) k1(approx) = f(t(R)(A),t(R)(B),w(h)(A),h(plateau), N) to yield the enantiomerization rate constant directly k(1). These values were compared with those obtained by computer-aided simulation with ChromWin. The Eyring activation parameters of the experimental interconversion profiles were determined to be: DeltaG(#)(298.15 K) = 103.6 +/- 0.9 kJ mol(-1), DeltaH(#) = 44.7 +/- 0.4 kJ mol(-1), DeltaS(#) = -198 +/- 7 J K(1) mol(-1) for dimethyl-1,3-allenedicarboxylate 1, and DeltaG(#)(298.15 K) = 103.5 +/- 1.1 kJ mol(-1), DeltaH(#) = 44.7 +/- 0.5 kJ mol(-1), DeltaS(#) = -197 +/- 9 J K(-1) mol(-1) for diethyl-1,3-allenedicarboxylate 2. The approximation function (AF) presented here allows the fast determination of rate constants k(1) and activation barriers of enantiomerization DeltaG(#) from chromatographic parameters without extensive computer simulation.  相似文献   

15.
The serpin conformational change by insertion of the reactive center loop into beta-sheet A plays a central role in multiple physiological consequences such as serine proteinase inhibition, latency and serpinopathic polymerization. To study the dynamic mechanism for the loop insertion, a novel kinetic method was established utilizing the ovalbumin mutant R339T/A352R; the loop insertion progressed after the cleavage of P1-P1' (Arg352-Ser353) by trypsin was quenched at pH 8 and 0.5 degrees C, and different conformers were quantified by separation using ion-exchange HPLC. The apparent first-order rate constant k(app) determined for various R339T/A352R derivatives differing in conformational stability was greatly increased by lowering the pH. The pH-dependence of k(app) indicated that the protonation of side-chain(s) with a pK(a) value of around 4.6 is a pre-requisite for the loop insertion. The theoretical rate constant k for the protonated form calculated from k(app) was highly variable, depending on the ovalbumin derivative; structural modifications that give increased mobility to helix F and the sheet-A half (s3A/s2A/s1A) resulted in a striking increase in the loop insertion rate constant k. The k values were determined at different temperatures for all the ovalbumin derivatives, and DeltaH(double dagger) and DeltaS(double dagger) values for the loop insertion reaction were determined according to the transition theory. The formation of the transition state was highly endothermic with minor entropy gain, requiring a DeltaG(double dagger) larger than 18 kcal/mol, which can offset the hydrogen-bond cleavages between s3A and s5A. These results are consistent with the transition state with an opened sheet A and altered orientation of helix F.  相似文献   

16.
Entropy-enthalpy (SH) compensation occurs when a small change in DeltaG is caused by large, and nearly compensatory, changes in DeltaH and DeltaS. It is considered a ubiquitous property of reactions in water. Because water is intimately involved in protein stability, SH compensation among protein variants, if it exists, could lead to important knowledge about protein-water interactions. In light of recent theoretical work on SH compensation, we gathered thermodynamic data for >200 protein variants to seek evidence for the simplest quantitative model of SH compensation (i.e., The van't Hoff denaturation enthalpy divided by the van't Hoff denaturation entropy is a constant). We conclude that either the data are insufficient to support the idea that quantitative SH compensation is a general feature of variant proteins or that such compensation does not exist. This study reinforces the idea that DeltaH-versus-DeltaS plots should not be used to provide evidence for SH compensation.  相似文献   

17.
Entropy was shown to play an equally important role as enthalpy for how enantioselectivity changes when redesigning an enzyme. By studying the temperature dependence of the enantiomeric ratio E of an enantioselective enzyme, its differential activation enthalpy (Delta(R-S)DeltaH(++)) and entropy (Delta(R-S)DeltaS(++)) components can be determined. This was done for the resolution of 3-methyl-2-butanol catalyzed by Candida antarctica lipase B and five variants with one or two point mutations. Delta(R-S)DeltaS(++) was in all cases equally significant as Delta(R-S)DeltaH(++) to E. One variant, T103G, displayed an increase in E, the others a decrease. The altered enantioselectivities of the variants were all related to simultaneous changes in Delta(R-S)DeltaH(++) and Delta(R-S)DeltaS(++). Although the changes in Delta(R-S)DeltaH(++) and Delta(R-S)DeltaS(++) were of a compensatory nature the compensation was not perfect, thereby allowing modifications of E. Both the W104H and the T103G variants displayed larger Delta(R-S)DeltaH(++) than wild type but exhibited a decrease or increase, respectively, in E due to their different relative increase in Delta(R-S)DeltaS(++).  相似文献   

18.
Trapp O 《Chirality》2006,18(7):489-497
An analytical solution of the unified equation to evaluate elution profiles of interconverting enantiomers in dynamic chromatography is presented. Rate constants k1 and k(-1) and Gibbs activation energies are directly obtained from the chromatographic parameters (retention times tR A and tR A of the interconverting enantiomers, the peak widths at half height wA and wB, and the relative plateau height hp), and the initial amounts A0 and B0 of the enantiomers without any iterative and time consuming computational step. Therefore, this equation is no longer limited to racemic analytes. The analytical solution presented here was validated by comparison with a dataset of 125,000 simulated elution profiles of enantiomerizations. Furthermore, it was found that the recovery rate from a defined dataset is on average 40% higher using the unified equation compared to evaluation methods based on iterative computer simulation. The new equation was applied to determine the enantiomerization rate constant of 1-n-butyl-2-tert-butyldiaziridine by enantioselective gas chromatography. The activation parameters (DeltaH(double dagger) = 112.6 +/- 2.5 kJ/mol and DeltaS(double dagger) = -27 +/- 2 J/(K mol) were obtained from temperature-dependent measurements between 100 degrees C and 140 degrees C in 10K steps.  相似文献   

19.
PYP (photoactive yellow protein) is a photoreceptor protein, which is activated upon photo-isomerization of the p-coumaric acid chromophore and is inactivated as the chromophore is thermally back-isomerized within a second (in PYP(M)-to-PYP(dark) conversion). Here we have examined the mechanism of the rapid thermal isomerization by analyzing mutant PYPs of Met100, which was previously shown to play a major role in facilitating the reaction [Devanathan, S. et al. (1998) Biochemistry 37, 11563-11568]. The mutation to Lys, Leu, Ala, or Glu decelerated the dark state recovery by one to three orders of magnitude. By evaluating temperature-dependence of the kinetics, it was found that the retardation resulted unequivocally from elevations of activation enthalpy (DeltaH( double dagger )) but not the other parameters such as activation entropy or heat capacity changes. Another effect exerted by the mutations was an up-shift of the apparent pK(a) of the chromophore [the pK(a) of a titratable group (X) that controls the pK(a) of the chromophore] in the PYP(M)-decay process. The pK(a) up-shift and the DeltaH( double dagger ) elevation show an approximately linear correlation. We, therefore, postulate that the role of Met100 is to reduce the energy barrier of the PYP(M)-decay process by an indirect interaction through X and that the process is thereby facilitated.  相似文献   

20.
Ethanol, isopropanol, propanol, and butanol enhanced thermal death in Saccharomyces cerevisiae by increasing DeltaSdouble dagger, the entropy of activation of thermal death while DeltaHdouble dagger, the enthalpy of activation, was not significantly affected. The relation between DeltaSdouble dagger and alkanol concentration was linear with a different slope for each alkanol: DeltaS(double dagger) (X) = DeltaS(double dagger) (0) + C(A) (E)X, where X is the alkanol concentration and C(A) (E) the entropy coefficient for the aqueous phase defined as increase in entropy of activation per unit concentrations of the alkanol. C(A) (E) was correlated with the lipid-buffer partition coefficients of the alkanols while C(M) (E), the entropy coefficient for the lipid phase, was nearly identical for the four alkanol and averaged 37.6 entropy units per mole of alkanol per kilogram of membrane. As predicted by these results, the specific death rates (K(d)) at constant temperature were an exponential function of the alkanol concentration and behaved in agreement with the following equation: In K(X) (d) = In K(0) (d) + (C(A) (E)/R)X, where R is the gas constant. It was concluded that the alkanols enhanced thermal death through nonspecific action on membrane structure.  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