首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Reversible flbrinogen polymer formation was examined at pH 6.6 and Γ/2 0.3. The equilibrium fraction of fibrinogen present as polymer, (Pmf)e, was determined by gel filtration for fibrinogen concentrations, FO, from 48 to 166 μm. Using FO in molarity, the experimental relation is ln [FO(Pmf)e] = 3.53 ln[FO(1 ? (Pmf)e)] + 23.73. This relation and attendant confidence limits are examined assuming, during filtration, that the original polymer population is either stable or selected polymer species dissociate to monomer. The possibility that all polymers are open is excluded since the calculated microscopic association constant would then increase with FO. Acceptable models are based on the assumptions that polymers are open, with association constant Ka, until restricted by closure, with association constant Kr, at an integral degree of polymerization, n. Values are selected on the basis that interaction parameters are independent of FO and that the required molar decrease in free energy is a minimum. Assuming polymer stability, the experimental relation at 273 °K gives n = 4, KrKa = 1.2 m, and Ka = 736 m?1. Temperature dependence gives ΔH= ?16.9 kcal/mol and ΔSOa = ?48.8 e.u. KrKa indicates a relation between changes in entropy. The probability is >0.90 that KrKa ? 56 m, which indicates a greater loss of degrees of freedom on closure than on association. Conclusions are not altered by the assumption that only the closed polymer species is stable. As ionic strength is decreased at pH 6.6, Ka increases. The clotting time of an otherwise constant system decreases as system Pmf is increased.  相似文献   

2.
The effects of pH on redox potentials of horseradish peroxidase-A and -(B + C) and of their heme-substituted enzymes with mesoheme, deuteroheme, chlorocruoroheme, and diacetyldeuteroheme were investigated. The slope in the plot of Eo against pK3 (a measure of basicity of pyrrole nitrogen) was found to be close upon 59 mVpK3 unit. It was also found that the ratio of ΔpKr to ΔpK3 was about 0.1 while that of ΔpKo to ΔpK3 was almost unity. Here, Kr and Ko stand for heme-linked proton dissociation constants in the ferrous and ferric peroxidases, respectively. The difference of either pKr or pKo between two isoenzyme preparations was about 1.6. These results support the previous conclusion (Arch. Biochem. Biophys. 165, 725, 1974 that Kr represents a proton dissociation constant of a distal amino acid residue and that there is a strong hydrogen bonding between its base and the water oxygen atom as a sixth ligand in the ferric state of peroxidases. The difference of redox potentials at pH 8.5 between two natural isoenzyme preparations, amounting to 88 mV, was attributed to the change in the hydrogen bonding strength caused by the difference in basicity of two distal amino acid residues. A possibility that approximate redox potentials of hemoproteins can be determined by analysis of several factors is discussed.  相似文献   

3.
Hydrolysis of benzyloxycarbonyl-GlyGlyPhe by nitro(Tyr 248)carboxypeptidase A over the pH range 4.88–8.04 has been examined. The nitroenzyme retains appreciable activity near pH 6.5, and the limiting value of Km is scarcely affected. The peptidase activity has a pH dependence characterized by the following parameters: pKE1 of 6.37 ± 0.19 and pKE2 of 6.60 ± 0.17 in kcatKm, and apparent pK of 5.59 ± 0.06 in Kcat. A spectroscopic pK of 6.75 ± 0.01, attributable to the nitro-Tyr 248 residue, has been determined. This correlates with the base-limb pKE2 in the kcatKm profile, which appears to be shifted from a higher value, pKE2 of 9.0, for the native enzyme. The single (acid-limb) pK which characterizes the kcat profile of the native enzyme is also found to be perturbed to a lesser extent by nitration. A kinetically competent reverse protonation mechanism, based on chemical modification and crystallographic evidence for the enzyme, is described.  相似文献   

4.
Using guanidinium and n-butylammonium cations (C+) as models for the positively charged side chains in arginine and lysine, we have determined the association constants with various oxyanions by potentiometric titration. For a dibasic acid, H2A, three association complexes may exist: K1M = [CHA][C+] [HA?]; K1D = [CA?][C+] [A2?]; K2D = [C2A][C+] [CA?]. For guanidinium ion and phosphate, K1M = 1.4, K1D = 2.6, and K2D = 5.1. The data for carboxylates indicate that the basicity of the oxyanion does not affect the association constant: acetate, pKa = 4.8, K1M = 0.37; formate, pKa = 3.8, K1M = 0.32; and chloroacetate, pKa = 2.9, K1M = 0.43, all with guanidinium ion. Association constants are also reported for carbonate, dimethylphosphinate, benzylphosphonate, and adenylate anions.  相似文献   

5.
6.
The effect of pH on the kinetic parameters for the chloroperoxidase-catalyzed N-demethylation of N,N-dimethylaniline supported by ethyl hydroperoxide was investigated from pH 3.0 to 7.0. Chloroperoxidase was found to be stable throughout the pH range studied. Initial rate conditions were determined throughout the pH range. The Vmax for the demethylation reaction exhibited a pH optimum at approximately 4.5. The Km for N,N-dimethylaniline increased with decreasing pH, while the Km for ethyl hydroperoxide varied in a manner paralleling Vmax. Comparison of the VmaxKm values for N,N-dimethylaniline and ethyl hydroperoxide indicated that the interaction of N,N-dimethylaniline with chloroperoxidase compound I was rate-limiting below pH 4.5, while compound I formation was rate-limiting above pH 4.5. The log of the VmaxKm for ethyl hydroperoxide was independent of pH, indicating that chloroperoxidase compound I formation is not affected by ionizations in this pH range. The plot of the log of the VmaxKm for N,N-dimethylaniline versus pH indicated an ionization on compound I with a pK of approximately 6.8. The plot of the log of the Vmax versus pH indicated an ionization on the compound I-N,N-dimethylaniline complex, with a pK of approximately 3.1. The results show that chloroperoxidase can demethylate both the protonated and neutral forms of N,N-dimethylaniline (pK approximately 5.0), suggesting that hydrophobic binding of the arylamine substrate is more important in catalysis than ionic bonding of the amine moiety. For optimal catalysis, a residue in the chloroperoxidase compound I-N,N-dimethylaniline complex with a pK of approximately 3.1 must be deprotonated, while a residue in compound I with a pK of approximately 6.8 must be protonated.  相似文献   

7.
The stoichiometry of free NADPH oxidation in phenobarbital induced rabbit liver microsomes was measured by means of registering the rates of NADPH, H+ and O2 consumption and O2? and H2O2 production. ΔO2?:ΔH2O2 ratio is approximately I indicating that about half H2O2 results from O2? dismutation, the second half being formed directly. ΔNADPH:ΔH2O2 and ΔO2:ΔH2O2 ratios exceed I and therefore another product of the reaction is water. The fact that the ratio (ΔNADPH-ΔH2O2):(ΔO2-ΔH2O2) is 2 allows one to consider direct 4-electron O2 reduction as the major way of water formation rather than endogenous substrate hydroxylation.  相似文献   

8.
The observed equilibrium constants (Kobs) for the l-phosphoserine phosphatase reaction [EC 3.1.3.3] have been determined under physiological conditions of temperature (38 °C) and ionic strength (0.25 m) and physiological ranges of pH and free [Mg2+]. Using Σ and square brackets to indicate total concentrations Kobs = Σ L-serine][Σ Pi]Σ L-phosphoserine]H2O], K = L-H · serine±]HPO42?][L-H · phosphoserine2?]H2O]. The value of Kobs has been found to be relatively sensitive to pH. At 38 °C, K+] = 0.2 m and free [Mg2+] = 0; Kobs = 80.6 m at pH 6.5, 52.7 m at pH 7.0 [ΔGobs0 = ?10.2 kJ/mol (?2.45 kcal/mol)], and 44.0 m at pH 8.0 ([H2O] = 1). The effect of the free [Mg2+] on Kobs was relatively slight; at pH 7.0 ([K+] = 0.2 m) Kobs = 52.0 m at free [Mg2+] = 10?3, m and 47.8 m at free [Mg2+] = 10?2, m. Kobs was insignificantly affected by variations in ionic strength (0.12–1.0 m) or temperature (4–43 °C) at pH 7.0. The value of K at 38 °C and I = 0.25 m has been calculated to be 34.2 ± 0.5 m [ΔGobs0 = ?9.12 kJ/mol (?2.18 kcal/ mol)]([H2O] = 1). The K for the phosphoserine phosphatase reaction has been combined with the K for the reaction of inorganic pyrophosphatase [EC 3.6.1.1] previously estimated under the same physiological conditions to calculate a value of 2.04 × 104, m [ΔGobs0 = ?28.0 kJ/mol (?6.69 kcal/mol)] for the K of the pyrophosphate:l-serine phosphotransferase [EC 2.7.1.80] reaction. Kobs = [Σ L-serine][Σ Pi][Σ L-phosphoserine][H2O], K = [L-H · serine±]HPO42?][L-H · phosphoserine2?]H2O. Values of Kobs for this reaction at 38 °C, pH 7.0, and I = 0.25 m are very sensitive to the free [Mg2+], being calculated to be 668 [ΔGobs0 = ?16.8 kJ/mol (?4.02 kcal/mol)] at free [Mg2+] = 0; 111 [ΔGobs0 = ?12.2 kJ/mol (?2.91 kcal/mol)] at free [Mg2+] = 10?3, m; and 9.1 [ΔGobs0 = ?5.7 kJ/mol (?1.4 kcal/mol) at free [Mg2+] = 10?2, m). Kobs for this reaction is also sensitive to pH. At pH 8.0 the corresponding values of Kobs are 4000 [ΔGobs0 = ?21.4 kJ/mol (?5.12 kcal/mol)] at free [Mg2+] = 0; and 97.4 [ΔGobs0 = ?11.8 kJ/ mol (?2.83 kcal/mol)] at free [Mg2+] = 10?3, m. Combining Kobs for the l-phosphoserine phosphatase reaction with Kobs for the reactions of d-3-phosphoglycerate dehydrogenase [EC 1.1.1.95] and l-phosphoserine aminotransferase [EC 2.6.1.52] previously determined under the same physiological conditions has allowed the calculation of Kobs for the overall biosynthesis of l-serine from d-3-phosphoglycerate. Kobs = [Σ L-serine][Σ NADH][Σ Pi][Σ α-ketoglutarate][Σ d-3-phosphoglycerate][Σ NAD+][Σ L-glutamat0] The value of Kobs for these combined reactions at 38 °C, pH 7.0, and I = 0.25 m (K+ as the monovalent cation) is 1.34 × 10?2, m at free [Mg2+] = 0 and 1.27 × 10?2, m at free [Mg2+] = 10?3, m.  相似文献   

9.
The pH dependence of the reaction of tris(hydroxymethyl)aminomethane (Tris) with the activated carbonyl compound 4-trans-benzylidene-2-phenyloxazolin-5-one (I) is given by the equation k′2 = kbKa(Ka + [H+]) + ka[OH?]Ka(Ka + [H+]), where Ka is the dissociation constant of TrisH+. Spectrophotometric experiments show that the Tris ester of α-benzamido-trans-cinnamic acid is formed quantitatively over a range of pH values, regardless of the relative contribution of kb and ka terms to k2. Hence, both terms refer to alcoholysis. While the mechanism of the reaction is not determined unequivocally in the present work, the magnitude of the kb term, together with its dependence on the basic form of Tris, suggests that ester formation is occurring by nucleophilic attack of a Tris hydroxyl group on the carbonyl carbon of the oxazolinone, with intramolecular catalysis by the Tris amino group. The rate enhancement due to this group is at least 102 and possibly of the order 106. This system is compared with other model systems for the acylation step of catalysis by serine esterases and proteinases.  相似文献   

10.
The rate of reaction of [Cr(III)Y]aq (Y is EDTA anion) with hydrogen peroxide was studied in aqueous nitrate media [μ = 0.10 M (KNO3)] at various temperatures. The general rate equation, Rate = k1 + k2K1[H+]?11 + K1[H+]?1 [Cr(III)Y]aq[H2O2] holds over the pH range 5–9. The decomposition reaction of H2O2 is believed to proceed via two pathways where both the aquo and hydroxo-quinquedentate EDTA complexes are acting as the catalyst centres. Substitution-controlled mechanisms are suggested and the values of the second-order rate constants k1 and k2 were found to be 1.75 × 10?2 M?1 s?1 and 0.174 M?1 s?1 at 303 K respectively, where k2 is the rate constant for the aquo species and k2 is that for the hydroxo complex. The respective activation enthalpies (ΔH*1 = 58.9 and ΔH*2 = 66.5 KJ mol?1) and activation entropies (ΔS*1 = ?85 and ΔS*2 = ?40 J mol?1 deg?1) were calculated from a least-squares fit to the Eyring plot. The ionisation constant pK1, was inferred from the kinetic data at 303 K to be 7.22. Beyond pH 9, the reaction is markedly retarded and ceases completely at pH ? 11. This inhibition was attributed in part to the continuous loss of the catalyst as a result of the simultaneous oxidation of Cr(III) to Cr(VI).  相似文献   

11.
12.
Four ethyl p-nitrophenyl alkylphosphonates were studied for the inhibition of elastase. A pH-dependence study using the assay substrate BOC-Ala-ONp or the phosphonate inhibitors showed the participation of an ionizing group with an apparent pKa of 6.9 and a decrease of reaction or inhibition at higher pH. Out of the four compounds investigated ethyl p-nitrophenyl pentylphosphonate was found to be the best inhibitor of elastase as judged from the value of k2KI. This value, which is the measure of inhibitory capacity, is the highest reported so far for the inhibition of elastase.  相似文献   

13.
Quenching of singlet molecular oxygen (1ΔgO2) by α-tocopherol (I) involves the hydroxy function of the chromanol ring of I. In phosphatidylcholine (PC) uni- and multilamellar vesicles this structural element of I is localized at the interface polar headgroup/hydrophobic core. A dielectric constant of ? ~ 25 was determined for this special region of the PC bilayer. The ratio kQ/kR of rate constants of quenching processes (kQ) and irreversible reactions (kR) of I with 1ΔgO2 increases with decreasing polarity of the solvent. In ethanolic solutions where ? = 25.5, kQ/kR is about 40. Extrapolation of these results to phospholipid bilayers suggests that at the nearness of the ester carbonyl oxygen of the PC fatty acid moieties, α-tocopherol can deactivate approximately 40 1ΔgO2 molecules before being destroyed. It is concluded that in vivo, one may expect to find a higher kQ/kR ratio if the chromanol ring of I hides within the more hydrophobic interiors of the membrane surface peptides.  相似文献   

14.
The electrical potential (Δψ) of intact cholinergic synaptic vesicles was measured in the presence and absence of the proton translocator carbonyl cyanide p-trifluoromethoxy-phenylhydrazone (FCCP), and the results were utilized to calculate the vesicular proton chemical gradient (ΔpH) and proton electrochemical potential μH+). At external pH = 7.4 the vesicles maintain a proton electrochemical gradient of ?+20 mV (positive inside) which is composed of Δψ??80 mV (negative inside) and ΔpH?1.6 (acidic inside). The proton chemical gradient (ΔpH) increases as a function of pHout whereas the vesicular electrical potential (Δψ) is only slightly affected by the external pH. Consequently, ΔμH+ is larger at basic external pH values (?+40 mV at pHout = 9.0) and smaller at acidic external pH values (ΔμH+?0 at (pHout = 5.6). The possible physiological role of the electrochemical potentials in maintaining high concentrations of acetylcholine within the cholinergic synaptic vesicle is discussed.  相似文献   

15.
The ionization of fatty acids, fatty amines and N-acylamino acids incorporated in phosphatidylcholine single-walled vesicles has been measured. The guest molecules have been specifically enriched with 13C and titrated by using NMR spectroscopy. The apparent pKa of fatty acids in phosphatidylcholine bilayers is 7.2–7.4 and those of fatty amines are approx. 9.5. These pKa values depend on many different parameters related to the structure of the lipid/ solution interface, to the composition of the aqueous medium and to the localization of the ionizable groups. A special sensitivity to the ionic strength and to the surface charge has been found. A positive surface charge decreases the pKa value whereas a negative one increases it, the total range of variation being 2.5–3 units. In a qualitative macroscopic interpretation, it is proposed that pKa is essentially determined by the low polarity of the lipidic matrix.  相似文献   

16.
A thermodynamic characterization of the Na+-H+ exchange system in Halobacterium halobium was carried out by evaluating the relevant phenomenological parameters derived from potential-jump measurements. The experiments were performed with sub-bacterial particles devoid of the purple membrane, in 1 M NaCl, 2 M KCl, and at pH 6.5–7.0. Jumps in either pH or pNa were brought about in the external medium, at zero electric potential difference across the membrane, and the resulting relaxation kinetics of protons and sodium flows were measured. It was found that the relaxation kinetics of the proton flow caused by a pH-jump follow a single exponential decay, and that the relaxation kinetics of both the proton and the sodium flows caused by a pNa-jump also follow single exponential decay patterns. In addition, it was found that the decay constants for the proton flow caused by a pH-jump and a pNa-jump have the same numerical value. The physical meaning of the decay constants has been elucidated in terms of the phenomenological coefficients (mobilities) and the buffering capacities of the system. The phenomenological coefficients for the Na+-H+ flows were determined as differential quantities. The value obtained for the total proton permeability through the particle membrane via all available channels, LH = (?JH +pH)Δψ,ΔpNa, was in the range of 850–1150 nmol H+·(mg protein)?1·h?1·(pH unit)?1 for four different preparations; for the total Na+ permeability, LNa = (?JNa+pNa)Δψ,ΔpH, it was 1620–2500 nmol Na+·(mg protein)?1·h?1·(pNa unit)?1; and for the proton ‘cross-permeability’, LHNa = (?JH+pNa)Δψ,ΔpH, it was 220–580 nmol H+·(mg protein)?1·h?1·(pNa unit)?1, for different preparations. From the above phenomenological parameters, the following quantities have been calculated: the degree of coupling (q), the maximal efficiency of Na+-H+ exchange (ηmax), the flow and force efficacies (?) of the above exchange, and the admissible range for the values of the molecular stoichiometry parameter (r). We found q ? 0.4; ηmax ? 5%; 0.36 ? r ? 2; ?JNa+ ? 1.3 · 105μmol · (RT unit)?1 at JNa = 1 μmolNa+ · (mgprotein)?1 · h?1; and ?ΔpNa ? 5 · 104 ΔpNa · (mg protein) · h · (RT unit)?1 at ΔpNa = 1 unit, for different preparations.  相似文献   

17.
Isolation and characterization of isocitrate lyase of castor endosperm   总被引:1,自引:0,他引:1  
Isocitrate lyase (threo-DS-isocitrate glyoxylate-lyase, EC 4.1.3.1) has been purified to homogeneity from castor endosperm. The enzyme is a tetrameric protein (molecular weight about 140,000; gel filtration) made up of apparently identical monomers (subunit molecular weight about 35,000; gel electrophoresis in the presence of sodium dodecyl sulfate). Thermal inactivation of purified enzyme at 40 and 45 °C shows a fast and a slow phase, each accounting for half of the intitial activity, consistent with the equation: At = A02 · e?k1t + A02 · e?k2t, where A0 and At are activities at time zero at t, and k1 and k2 are first-order rate constants for the fast and slow phases, respectively. The enzyme shows optimum activity at pH 7.2–7.3. Effect of [S]on enzyme activity at different pH values (6.0–7.5) suggests that the proton behaves formally as an “uncompetitive inhibitor.” A basic group of the enzyme (site) is protonated in this pH range in the presence of substrate only, with a pKa equal to 6.9. Successive dialysis against EDTA and phosphate buffer, pH 7.0, at 0 °C gives an enzymatically inactive protein. This protein shows kinetics of thermal inactivation identical to the untreated (native) enzyme. Full activity is restored on adding Mg2+ (5.0 mm) to a solution of this protein. Addition of Ba2+ or Mn2+ brings about partial recovery. Other metal ions are not effective.  相似文献   

18.
The reactivities of anionic nitroalkanes with 2-nitropropane dioxygenase of Hansenula mrakii, glucose oxidase of Aspergillus niger, and mammalian d-amino acid oxidase have been compared kinetically. 2-Nitropropane dioxygenase is 1200 and 4800 times more active with anionic 2-nitropropane than d-amino acid oxidase and glucose oxidase, respectively. The apparent Km values for anionic 2-nitropropane are as follows: 2-nitropropane dioxygenase, 1.61 mm; glucose oxidase, 16.7 mm; and d-amino acid oxidase, 11.1 mm. Anionic 2-nitropropane undergoes an oxygenase reaction with 2-nitropropane dioxygenase and glucose oxidase, and an oxidase reaction with d-amino acid oxidase. In contrast, anionic nitroethane is oxidized through an oxygenase reaction by 2-nitropropane dioxygenase, and through an oxidase reaction by glucose oxidase. All nitroalkane oxidations by these three flavoenzymes are inhibited by Cu and Zn-superoxide dismutase of bovine blood, Mn-superoxide dismutases of bacilli, Fe-superoxide dismutase of Serratia marcescens, and other O2? scavengers such as cytochrome c and NADH, but are not affected by hydroxyl radical scavengers such as mannitol. None of the O2? scavengers tested affected the inherent substrate oxidation by glucose oxidase and d-amino acid oxidase. Furthermore, the generation of O2? in the oxidation of anionic 2-nitropropane by 2-nitropropane dioxygenase was revealed by ESR spectroscoy. The ESR spectrum of anionic 2-nitropropane plus 2-nitropropane dioxygenase shows signals at g1 = 2.007 and g11 = 2.051, which are characteristic of O2?. The O2? generated is a catalytically essential intermediate in the oxidation of anionic nitroalkanes by the enzymes.  相似文献   

19.
Proton inventory investigations of the hydrolysis N-acetylbenzotriazole at pH 3.0 (or the equivalent point on the pD rate profile) have been conducted at two different temperatures and at ionic strengths ranging from 0 to 3.0 M. The solvent deuterium isotope effects and proton inventories are remarkably similar over this wide range of conditions. The proton inventories suggest a cyclic transition state involving four protons contributing to the solvent deuterium isotope effect for the water-catalyzed hydrolysis. The hydrolysis data are described by the equation kn = ko (1 ? n + nπa1)4 with πa1 ~ 0.74, where ko is the observed first-order rate constant in protium oxide, n is the atom fraction of deuterium in the solvent, kn is the rate constant in a protium oxide-deuterium oxide mixture, and πa1 is the isotopic fractionation factor.  相似文献   

20.
The magnesium ion-dependent equilibrium of vacant ribosome couples with their subunits
70 S?k?1k150 S+30S
has been studied quantitatively with a novel equilibrium displacement labeling method which is more sensitive and precise than light-scattering. At a concentration of 10?7m, tight couples (ribosomes most active in protein synthesis) dissociate between 1 and 3 mm-Mg2+ at 37 °C with a 50% point at 1.9 mm. The corresponding association constants Ka′ are 5.1 × 105m?1 (1 mm-Mg2+), 3.5 × 107m?1 (2 mm), and 1.2 × 109m?1 (3 mm), about five orders of magnitude higher than the Ka′ value of loose couples studied by Spirin et al. (1971) and Zitomer & Flaks (1972).In this range of Mg2+ concentrations (37 °C, 50 mm-NH4+) the rate constants depend exponentially and in opposite ways on the Mg2+ concentration: k1 = 2.2 × 10?3s?1, k?1 = 7.7 × 104m?1s?1 (2mm-Mg2+); k1 = 1.5 × 10?4s?1, k?1 = 1.7 × 107m?1s?1 (5 mm-Mg2+). Under physiological conditions (Mg2+ ~- 4 mm, ribosome concn ~- 10?7m), the equilibrium strongly favors association and the rate of exchange is slow (t12 ~- 10 min). In the range of dissociation (2 mm-Mg2+), association of subunits proceeds without measurable entropy change and hence ΔGO = ΔHO. The negative enthalpy change of ΔHO = ? 10 kcal suggests that association of subunits involves a shape change.Below a critical Mg2+ concentration (~- 2 mm), the 50 S subunits are converted irreversibly into the b-form responsible for the transition to loose couples. The results are compatible with two classes of binding sites, one class binding Mg2+ non-co-operatively and contributing to the free energy of association by reduction of electrostatic repulsion, and another class probably consisting of hydrogen bonds between components at opposite interfaces whose critical spatial alignment rapidly denatures in the absence of stabilizing magnesium ions.  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