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 46 毫秒
1.
Inactivation of alpha- and beta-thrombin by alpha 2-macroglobulin, by alpha 1-proteinase inhibitor and by antithrombin-III and heparin was studied. The amount of alpha- and beta-thrombin inactivated by antithrombin-III was proportional to the concentration of the inhibitor, but the inactivation rates of the two forms of thrombin were different. Heparin facilitated complex-formation between alpha-thrombin and antithrombin-III, whereas inactivation of beta-thrombin by antithrombin was only slightly influenced, even at a heparin concentration two orders of magnitude higher. alpha 2-Macroglobulin inhibited both alpha- and beta-thrombin activity similarly, i.e. the amount of alpha- and beta-thrombin inactivated as well as the rates of their inhibition were the same. alpha 1-Proteinase inhibitor also formed a complex with alpha- and beta-thrombin, similarly to antithrombin-III, although the inactivation of the enzyme needed high inhibitor concentration and long incubation time. These results suggest that the inactivation of beta-thrombin, if it occurs in the plasma, is also controlled by plasma inhibitors.  相似文献   

2.
The influence of PGI2 on the activity and on the inactivation of enzymes participating in blood coagulation /thrombin and Factor Xa/ and fibrinolysis /plasmin/ were investigated. According to the results PGI2 has no effect on the activity of Factor Xa and plasmin nor on the inactivation of these enzymes by antithrombin-III in the absence and presence of heparin at a concentration of PGI2 up to 400 μg/ml. An acceleration of the inactivation of thrombin by antithrombin-III was found in the presence of PGI2 within a concentration of 100–400 μg/ml without any effect on the heparin-accelerated inactivation of thrombin by antithrombin. We got similar results using clotting tests for the assay and the application of synthetic substrate for thrombin. This inactivation-accelerating effect of PGI2 on thrombin was only demonstratable at a concentration five magnitudes higher than that of the anti-aggregation effect on platelets.  相似文献   

3.
Inactivation of plasmin by a 3-4-fold molar excess of antithrombin III follows pseudo-first-order kinetics and the apparent rate constants are proportional to the concentration of the inhibitor. Heparin accelerates the inactivation reaction without changing its pseudo-first-order character, and the apparent rate constants are also proportional to the concentration of the polysaccharide. Heparin results in a minimum 20-fold rate enhancement of the reaction between plasmin and antithrombin III when the concentrations of heparin and plasmin are approx. 0.5mum and 1mum respectively. Heparin at a molar concentration well below that of plasmin still accelerates the reaction: one molecule of the polysaccharide is able to facilitate the inactivation of about 100 molecules of plasmin. Heparin must bind to plasmin to accelerate the plasmin-antithrombin III reaction, since the modification of four to five lysine residues of the enzyme inhibits the rate-enhancement effect of heparin and the dissociation of heparin-plasmin complex decreases the inactivation rate of plasmin. Increasing the concentration of antithrombin III, at a constant amount of heparin, results in increase of the inactivation rate. By contrast, the effect of increasing the amount of plasmin in the presence of constant amount of heparin and antithrombin III is such that higher plasmin-to-heparin ratios are associated with lower rates of inactivation. It seems, therefore, that to obtain ;optimal' conditions for fast enzyme inactivation, the amount of heparin should be matched to plasmin rather than to antithrombin III. Arrhenius plots of the plasmin-antithrombin III reaction are linear both in the absence and presence of heparin, at concentrations of 1 or 2mug/ml, over a range of 26K. Under these experimental conditions, heparin increases activation entropy. The findings show that heparin seems to fulfil some criteria that are characteristic for biological catalysis: binding, reaction-rate enhancement (increasing activation entropy), recycling of heparin (effectiveness of non-stoichiometric amounts of the polysaccharide) and specificity.  相似文献   

4.
The inactivation of thrombin by serum follows a first order kinetics with thrombin concentrations of the order of 20 to 40 units per ml. of serum. With higher thrombin concentrations of 300 to 600 units per ml. of serum, the rate slows down and not the logarithm of the clotting time, but the clotting time itself increases proportionally with the incubation time. The antithrombic factor is stable at +4°C. for a period of 2 weeks, or for 3 months at –15°C. Heating over 58°C. destroys it rapidly. The heat inactivation has an energy of activation of 121,000 cal. With low thrombin concentrations the rate of thrombin inactivation increases linearly with increasing serum concentration. The pH optimum of the inactivation is at 8.5. Increasing the temperature increases the rate. The energy of activation of the thrombin destruction by serum is 14,000 cal. Heparin increases the rate considerably with low thrombin concentrations, but does not affect the rate with high thrombin concentrations. The effect of heparin can be abolished with protamine or toluidine blue. Some other reagents were also tested with respect to their effect upon the rate of inactivation of thrombin by serum.  相似文献   

5.
The influence of PGI2 on the activity and on the inactivation of enzymes participating in blood coagulation (thrombin and Factor Xa) and fibrinolysis (plasmin) were investigated. According to the results PGI2 has no effect on the activity of Factor Xa and plasmin nor on the inactivation of these enzymes by antithrombin-III in the absence and presence of heparin at a concentration of PGI2 up to 400 micrograms/ml. An acceleration of the inactivation of thrombin by antithormbin-III was found in the presence of PGI2 within a concentration of 100-400 micrograms/ml without any effect on the heparin-accelerated inactivation of thrombin by antithrombin. We got similar results using clotting tests for the assay and the application of synthetic substrate for thrombin. This inactivation-accelerating effect of PGI2 on thrombin was only demonstratable at a concentration five magnitudes higher than that of the anti-aggregation effect on platelets.  相似文献   

6.
The heat of the reaction NAD(+)+propan-2-ol=NADH+acetone+H(+) was determined to be 42.5+/-0.6kJ/mol (10.17+/-0.15kcal/mol) from equilibrium measurements at 9-42 degrees C catalysed by yeast alcohol dehydrogenase. With the aid of thermochemical data for acetone and propan-2-ol the values of DeltaH=-29.2kJ/mol (-6.99kcal/mol) and DeltaG(0)=22.1kJ/mol (5.28kcal/mol) are derived for the reduction of NAD (NAD(+)+H(2)=NADH+H(+)). These values are consistent with analogous but less accurate data for the ethanol-acetaldehyde reaction. Thermodynamic data for the reduction of NAD and NADP are summarized.  相似文献   

7.
Enzymatic and thermodynamic characteristics of type II isopentenyl diphosphate (IPP):dimethylallyl diphosphate (DMAPP) isomerase (Tk-IDI) from Thermococcus kodakaraensis, which catalyzes the interconversion of IPP and DMAPP, were examined. FMN was tightly bound to Tk-IDI, and the enzyme required NADPH and Mg2+ for the isomerization in both directions. The melting temperature (Tm), the change of enthalpy (deltaH(m)), and the heat capacity change (deltaC(p)) of Tk-IDI were 88.0 degrees C, 444 kJ mol(-1), and 13.2 kJ mol(-1) K(-1), respectively, indicating that Tk-IDI is fairly thermostable. Kinetic parameters dramatically changed when the temperature crossed 80 degrees C even though its native overall structure was stably maintained up to 90 degrees C, suggesting that local conformational change would occur around 80 degrees C. This speculation was supported by the result of the circular dichroism analysis that showed the shift of the alpha-helical content occurred at 80 degrees C.  相似文献   

8.
Binding of heparin to human high molecular weight kininogen   总被引:1,自引:0,他引:1  
The binding of heparin to high molecular weight kininogen (H-kininogen) was analyzed by the effect of kininogen in decreasing the heparin-induced enhancement of the rate of inactivation of thrombin by antithrombin. The conditions were arranged so that the heparin-catalyzed antithrombin-thrombin reaction, monitored in the presence of the reversible thrombin inhibitor p-aminobenzamidine, followed pseudo-first-order kinetics and the observed rate constant (kappa obsd) varied linearly with the heparin concentration. In the absence of metal ions, H-kininogen minimally affected kappa obsd, measured at a constant concentration of heparin with high affinity for antithrombin (30 nM), at I = 0.15, pH 7.4 and 25 degrees C. However, at a saturating concentration of Zn2+ (10 microM), kappa obsd was reduced to 50% at approximately 20 nM H-kininogen and to that of the uncatalyzed reaction at greater than or equal to approximately 0.2 microM H-kininogen. Conversely, at a saturating concentration of H-kininogen (0.5 microM), kappa obsd was decreased to 50% at approximately 0.6 microM Zn2+ and to the kappa obsd of the uncatalyzed reaction at greater than or equal to 10 microM Zn2+. Other metal ions were effective in the order Zn2+ approximately Ni2+ greater than Cu2+ approximately Co2+ approximately Cd2+. The single-chain and two-chain forms of H-kininogen and the H-kininogen light chain reduced the heparin enhancement in the presence of Zn2+ to the same extent, whereas low molecular weight kininogen had no influence.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The kinetics of cyanide binding to chloroperoxidase were studied using a high-pressure stopped-flow technique at 25 degrees C and pH 4.7 in a pressure range from 1 to 1000 bar. The activation volume change for the association reaction is delta V not equal to + = -2.5 +/- 0.5 ml/mol. The total reaction volume change, determined from the pressure dependence of the equilibrium constant, is delta V degrees = -17.8 +/- 1.3 ml/mol. The effect of temperature was studied at 1 bar yielding delta H not equal to + = 29 +/- 1 kJ/mol, delta S not equal to + = -58 +/- 4 J/mol per K. Equilibrium studies give delta H degrees = -41 +/- 3 kJ/mol and delta S degrees = -59 +/- 10 J/mol per K. Possible contributions to the binding process are discussed: changes in spin state, bond formation and conformation changes in the protein. An activation volume analog of the Hammond postulate is considered.  相似文献   

10.
(-)-Menthide, a seven-membered lactone derived from the natural product (-)-menthol, was polymerized using a structurally defined zinc-alkoxide catalyst to form an aliphatic polyester. The polymer was fully characterized by NMR spectroscopy, size exclusion chromatography, and matrix-assisted laser desorption ionization mass spectrometry. The polymerization reaction occurred in a controlled fashion and polymer samples with M(n) values up to 90 kg mol(-1) were obtained by varying the catalyst loading. Monitoring of the rate of polymerization by in situ FT-IR spectroscopy (ReactIR) revealed a first order dependence on (-)-menthide. The temperature dependence of the observed rate constant between 30 and 90 degrees C was well described by the Arrhenius equation and gave E(a) = 38.4 +/- 0.9 kJ mol(-1). Thermodynamic parameters (deltaH(p) degrees = -16.8 +/- 1.6 kJ mol(-1), deltaS(p) degrees = -27.4 +/- 4.6 J mol(-1) K(-1)) for the polymerization of (-)-menthide were also determined by measuring the equilibrium monomer concentration at different temperatures ranging from 40 to 100 degrees C. The equilibrium monomer concentrations at 25 and 100 degrees C were calculated to be 0.031 +/- 0.018 and 0.120 +/- 0.063 M, respectively.  相似文献   

11.
A calorimetric study has been made of the heats of CO reaction with the monomeric haemoglobins of Chironomus thummi thummi III and IV as a function of pH. The number of Bohr protons released at pH 7.1 was determined from heats of reaction in different buffers as 0.19 and 0.31 mol H+/mol CO for haemoglobin III and IV respectively. The heat of the Bohr ionization process was found to be 6 and 8 kcal/mol H+ (25 and 34 kJ/mol) for the haemoglobins III and IV. These values are consistent with values found for histidine groups. A pH-independent part of the reaction enthalpy was determined as - 19.7 kcal/mol CO (-82.4 kJ/mol). The same reaction with myoglobin is less exothermic. From the combination of deltaG0 and deltaH0 values TdeltaS0 values have been calculated. It was found for both haemoglobins that the entropy of reaction is greater by 2 cal K-1 mol-1 (8.4 JK-1 mol-1) at pH 9.5 as compared to pH 6.0.  相似文献   

12.
The inhibition of thrombin by antithrombin III (AT III) and heparin has been studied in pure systems to determine the kinetics of inhibition during human prothrombin activation. The present study shows that prothrombinase-catalyzed prothrombin activation resulted in the generation of thrombin and meizothrombin(des F1). In the absence of heparin the second-order rate constants of the inactivation of both thrombin and meizothrombin(des F1) formed in the reaction mixture appeared to be identical, k = 3.7 X 10(5) M-1 min-1. The rate constant of inhibition of purified thrombin was 6.5 X 10(5) M-1 min-1. In the presence of heparin the decay of the amidolytic activity was biexponential and could be modeled by a four-parameter equation to determine the pseudo first-order rate constants of inhibition as well as the composition of the reaction with respect to the levels of thrombin and meizothrombin(des F1). The ratio of thrombin over meizothrombin(des F1) varied with the initial prothrombin concentration. Heparin catalyzed the AT III inhibition of thrombin but not meizothrombin(des F1) formed during the prothrombin activation. Thrombin, generated by (Xa-Va-phospholipid-Ca2+) was inhibited by AT III/heparin more slowly than purified thrombin, and the saturation kinetics of the inhibition with respect to AT III differed from those found with purified thrombin.  相似文献   

13.
The thermodynamics of the conversion of aqueous L-aspartic acid to fumaric acid and ammonia have been investigated using both heat conduction microcalorimetry and high-pressure liquid chromatography. The reaction was carried out in aqueous phosphate buffer over the pH range 7.25-7.43, the temperature range 13-43 degrees C, and at ionic strengths varying from 0.066 to 0.366 mol kg(-1). The following values have been found for the conversion of aqueous L-aspartateH- to fumarate2- and NH4+ at 25 degrees C and at zero ionic strength: K = (1.48 +/- 0.10) x 10(-3), DeltaG degrees = 16.15 +/- 0.16 kJ mol(-1), DeltaH degrees = 24.5 +/- 1.0 kJ mol(-1), and DeltaC(p) degrees = -147 +/- 100 J mol(-1) K(-1). Calculations have also been performed which give values of the apparent equilibrium constant for the conversion of L-aspartic acid to fumaric acid and ammonia as a function of temperature, pH and ionic strength.  相似文献   

14.
A calorimetric titration method was used to study ADP binding to native myosin. Data were analyzed by assuming that the myosin molecule has n independent and identical sites for ADP binding. The enthalpy change (deltaH), the binding constant (K), and n were determined. In 0.5 M KCl, 0.01 M MgCl2, and 0.02 M Tris/HCl (pH 7.8), we found: at 0 degrees, deltaH = -57.1 +/- 3.2 kJ-mol-1, log K = 6.42 +/- 0.13, n = 1.49 +/- 0.07; at 12 degrees, deltaH = 73.1 +/- 3.2 kJ-mole-1, log K = 6.08 +/- 0.13, and n = 1.74 +/- 0.07. The average heat capacity change on ADP binding to myosin between 0 and 12 degrees is thus -1.4 +/- 0.4 kJ-mol-1-K-1. Reasonably consistent results were obtained at 25 degrees, suggesting ADP binding to myosin is as strongly exothermic as at lower temperatures, although further interpretation of this result seems unwarranted, mainly because of the instability of myosic at this temperature. The number of protons released on binding of ADP to myosin was determined in separate experiments. The value was 0.19 +/- 0.02 at both 0 and 12 degrees. The reaction of protons with Tris thus contributes about -9.5 kJ-mol-1 to the observed heat on ADP binding.  相似文献   

15.
Dihydroorotase (DHOase, EC 3.5.2.3) from the extreme thermophile Bacillus caldolyticus has been subcloned, sequenced, expressed, and purified as a monomer. The catalytic properties of this thermophilic DHOase have been compared with another type I enzyme, the DHOase domain from hamster, to investigate how the thermophilic enzyme is adapted to higher temperatures. B. caldolyticus DHOase has higher Vmax and Ks values than hamster DHOase at the same temperature. The thermodynamic parameters for the binding of L-dihydroorotate were determined at 25 degrees C for hamster DHOase (deltaG = -6.9 kcal/mol, deltaH = -11.5 kcal/mol, TdeltaS = -4.6 kcal/mol) and B. caldolyticus DHOase (deltaG = -5.6 kcal/mol, deltaH = -4.2 kcal/mol, TdeltaS = +1.4 kcal/mol). The smaller enthalpy release and positive entropy for thermophilic DHOase are indicative of a weakly interacting Michaelis complex. Hamster DHOase has an enthalpy of activation of 12.3 kcal/mol, similar to the release of enthalpy upon substrate binding, rendering the kcat/Ks value almost temperature independent. B. caldolyticus DHOase shows a decrease in the enthalpy of activation from 12.2 kcal/mol at temperatures from 30 to 50 degrees C to 5.3 kcal/mol for temperatures of 50-70 degrees C. Vibrational energy at higher temperatures may facilitate the transition ES --> ES(double dagger), making kcat/Ks almost temperature independent. The pseudo-first-order rate constant for water attack on L-dihydroorotate, based on experiments at elevated temperature, is 3.2 x 10(-11) s(-1) at 25 degrees C, with deltaH(double dagger) = 24.7 kcal/mol and TdeltaS(double dagger) = -6.9 kcal/mol. Thus, hamster DHOase enhances the rate of substrate hydrolysis by a factor of 1.6 x 10(14), achieving this rate enhancement almost entirely by lowering the enthalpy of activation (delta deltaH(double dagger) = -19.5 kcal/mol). Both the rate enhancement and transition state affinity of hamster DHOase increase steeply with decreasing temperature, consistent with the development of H-bonds and electrostatic interactions in the transition state that were not present in the enzyme-substrate complex in the ground state.  相似文献   

16.
The thermodynamics of 5'-ATGCTGATGC-3' binding to its complementary DNA and RNA strands was determined in sodium phosphate buffer under varying conditions of temperature and salt concentration from isothermal titration calorimetry (ITC). The Gibbs free energy change, DeltaG degrees of the DNA hybridization reactions increased by about 6 kJ mol(-1) from 20 degrees C to 37 degrees C and exhibited heat capacity changes of -1.42 +/- 0.09 kJ mol(-1) K(-1) for DNA/DNA and -0.87 +/- 0.05 kJ mol(-1) K(-1) for DNA/RNA. Values of DeltaG degrees decreased non-linearly by 3.5 kJ mol(-1) at 25 degrees C and 6.0 kJ mol(-1) at 37 degrees C with increase in the log of the sodium chloride concentration from 0.10 M to 1.0 M. A near-linear relationship was observed, however, between DeltaG degrees and the activity coefficient of the water component of the salt solutions. The thermodynamic parameters of the hybridization reaction along with the heat capacity changes were combined with thermodynamic contributions from the stacking to unstacking transitions of the single-stranded oligonucleotides from differential scanning calorimetry (DSC) measurements, resulting in good agreement with extrapolation of the free energy changes to 37 degrees C from the melting transition at 56 degrees C.  相似文献   

17.
We study the heat associated with the transformation of vesicles into micelles in mixtures of bilayer-forming phospholipids and micelle-forming surfactants. We subdivide the total heat evolution deltaQ(coex) within the range of coexistence of vesicles and micelles into three contributions related to the transition of dN(D)m-b molecules of surfactant and dN(L)m-b molecules of lipid from micelles to vesicles and to the extraction of dN(D)m-w molecules of surfactant from micelles to the aqueous solution, so that deltaQ(coex) = deltaH(D)m-w x dN(D)m-w + deltaH(D)m-b x dN(D)m-b + deltaH(L)m-b x dN(L)m-b where deltaH(D)m-w, deltaH(L)m-b, and deltaH(D)m-b are the respective molar "transfer" enthalpies. We design a method for the evaluation of all three molar enthalpies, from isothermal calorimetric titrations conducted according to two different protocols of titration of lipid-surfactant mixtures. In the first protocol the mixture is titrated with an aqueous solution of pure lipid vesicles, and in the second the mixture is titrated with an aqueous solution of pure surfactant. Titration of the mixed systems by a buffer solution serves to verify the results obtained under these protocols. In addition to the values of molar enthalpies, our method yields the cmc value of the pure surfactant. We apply our method to investigating the heat evolution in mixtures of egg yolk phosphatidylcholine and the nonionic surfactant octylglucoside in a phosphate-buffered saline solution at 28 degrees C. These studies gave the following values: deltaH(D)m-w = -1732 cal/mol, deltaH(L)m-b = -592 cal/mol, deltaH(D)m-b = 645 cal/mol, and cmc = 23.5 mM. We discuss the possible physical insight of these values and the perspectives of applications of the proposed method.  相似文献   

18.
Heparin inhibits (I50 = 2 microgram/ml) the activity of luteinizing hormone and human chorionic gonadotropin-stimulated adenylate cyclase in purified rat ovarian plasma membranes. Unstimulated enzyme activity and activity stimulated by NaF, GTP or guanosine 5'-(beta,gamma-imido)triphosphate were inhibited to a lesser extent. Human chorionic gonadotropin binding to this membrane preparation was inhibited by heparin (I50 = 6 microgram/ml). The inhibition with respect to hormone concentration was of a mixed type for hormone binding and adenylate cyclase stimulation. Inhibition by heparin was not eliminated at saturating hormone concentration. The degree of inhibition was unaffected by the order in which enzyme, hormone and heparin were introduced into the assay system. Heparin (3 microgram/ml) did not affect the pH activity relationship of basal and hormone-stimulated adenylate cyclase activity and did not change the dependence of enzyme activity on magnesium ion concentration. The inhibitory action of heparin cannot be solely attributed to interference with either catalysis or hormone binding. The possibility is considered that the highly charged heparin molecule interferes with enzyme receptor coupling, by restricting the mobility of these components or by effecting their conformation.  相似文献   

19.
The thermodynamics of binding of the antibacterial peptide magainin 2 amide (M2a) to negatively charged small (SUVs) and large (LUVs) unilamellar vesicles has been studied with isothermal titration calorimetry (ITC) and CD spectroscopy at 45 degrees C. The binding isotherms as well as the ability of the peptide to permeabilize membranes were found to be qualitatively and quantitatively similar for both model membranes. The binding isotherms could be described with a surface partition equilibrium where the surface concentration of the peptide immediately above the plane of binding was calculated with the Gouy-Chapman theory. The standard free energy of binding was deltaG0 approximately -22 kJ/mol and was almost identical for LUVs and SUVs. However, the standard enthalpy and entropy of binding were distinctly higher for LUVs (deltaH0 = -15.1 kJ/mol, deltaS0 = 24.7 J/molK) than for SUVs (deltaH0 = -38.5 kJ/mol, deltaS0 = -55.3 J/molK). This enthalpy-entropy compensation mechanism is explained by differences in the lipid packing. The cohesive forces between lipid molecules are larger in well-packed LUVs and incorporation of M2a leads to a stronger disruption of cohesive forces and to a larger increase in the lipid flexibility than peptide incorporation into the more disordered SUVs. At 45 degrees C the peptide easily translocates from the outer to the inner monolayer as judged from the simulation of the ITC curves.  相似文献   

20.
A calorimetric titration method was used to study the ADP binding to the chymotryptic subfragments of myosin, heavy meromyosin (HMM) and myosin subfragment 1 (S-1), and to myosin aggregated into filaments at low ionic strength. The binding constant (K) and heat of reaction (deltaH, kiloJoules (moles of ADP bound)-1) were determined. For HMM in 0.5 M KCl, 0.01 M MgCl2, 0.02 M Tris (pH 7.8) at 12 degrees, log K = 5.92 +/- 0.13 and deltaH = -70.9 +/- 3.6 kJ mol-1. These results agree with our previous findings for myosin in 0.5 M KCl at 12 degrees. When the KCl concentration was reduced to 0.1 M, the binding constant did not change significantly (log K = 6.09 +/- 0.06) but the binding was more exothermic (deltaH = -90.1 +/- 3.3 kJ mol-1). Similar results were obtained for myosin filaments in 0.1 M KCl and also for both the isoenzymes of S-1(S-1(A1) and S-1(A2) in 0.1 M KCl. In 0.5 M KCl, the binding curves suggest that about one ADP is bound per active site, but as 0.1 M KCl, the apparent stoichiometry drops from 0.7 to 0.75. The most probable explanation is that there is some site heterogeneity which is more evident at lower ionic strength.  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