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.
The degradation kinetics of aplidine were investigated using reversed-phase high-performance liquid chromatography combined with UV detection. Aplidine consists of at least two isomers that undergo interconversion at a low rate. Influences of pH, temperature, buffer ions and ionic strength on the degradation kinetics were studied. The log kobs–pH profile can be divided into three parts, a proton, a solvent and a hydroxyl-catalysed section. The stability-indicating properties of the used analysis technique as well as the identities of the main degradation products were checked using gradient liquid chromatography and mass spectrometric detection. The overall degradation rate constant as a function of the temperature under acidic and alkaline conditions obeys the Arrhenius equation. No catalytic influences were observed with phosphate and carbonate buffers and, in addition, the ionic strength showed no substantial effect on the stability, as expected. Results from gradient LC–MS indicated that hydrolysis of the ester groups present in the ring structure was the main degradation route. There is no difference in degradation rate constants for the individual isomers.  相似文献   

2.
Hyaluronan (HA) hydrolysis catalyzed by hyaluronidase (HAase) is inhibited at low HAase over HA ratio and low ionic strength, because HA forms electrostatic complexes with HAase, which is unable to catalyze hydrolysis. Bovine serum albumin (BSA) was used as a model to study the HA-protein electrostatic complexes at pH 4. At low ionic strength, there is formation of (i) neutral insoluble complexes at the phase separation and (ii) small positively-charged or large negatively-charged soluble complexes whether BSA or HA is in excess. According to the ionic strength, different types of complex are formed. Assays for HA and BSA led to the determination of the stoichiometry of these complexes. HAase was also shown to form the various types of complex with HA at low ionic strength. Finally, we showed that at 0 and 150 mmol L(-1) NaCl, BSA competes with HAase in forming complexes with HA and thus induces HAase release resulting in a large increase in the hydrolysis rate. These results, in addition to data in the literature, show that HA-protein complexes, which can exist under numerous and varied conditions of pH, ionic strength and protein over HA ratio, might control the in vivo HAase activity.  相似文献   

3.
The Dictyostelium discoideum alkaline phosphatase was investigated kinetically in an attempt to elucidate its mechanism of action. Analysis of the hydrolysis of p-nitrophenyl phosphate by stopped-flow spectrophotometry revealed biphasic kinetics, suggesting a double displacement enzyme mechanism. Furthermore, Tris stimulated activity in an uncompetitive manner, a result that was consistent with this interpretation. The enzyme was inhibited reversibly by phosphate at low ionic strength, but the inhibition was irreversible at high ionic strength and the latter effect was enhanced at alkaline pH values. These results indicate that high ionic strength and alkaline pH conditions bring about a conformational change that renders the enzyme susceptible to irreversible inhibition by phosphate.  相似文献   

4.
Initiation factor IF-2 ribosome dependent GTP hydrolysis (uncoupled GTPase) presents a bell-shaped pH profile which is shifted by changes in ionic strength. At low ionic strength (I = 25 mM) the maximal hydrolytic activity occurs at pH 7.5; when the ionic strength is increased the pH optimum of the reaction is shifted toward more acidic values. Such behavior can be satisfactorily explained as the effect of an electrostatic potential developed by a neighboring polyanion, presumably RNA, on the catalytic site. The addition of fMet-tRNAfMet or AcPhe-tRNAPhe and messenger RNA (coupled GTPase) changes the ionic strength--pH characteristics of the reaction. Thus there is an effect, direct or indirect, of components located at the ribosomal P site. Investigation of the effect of neighboring polyanions on the catalytic activity of the factor-dependent ribosomal GTPases can be seen to provide information about their functional significance that is complementary to that gained from direct structural studies.  相似文献   

5.
The passive permeability of the red blood cell in cations   总被引:6,自引:3,他引:3       下载免费PDF全文
The efflux of salt from human red blood cells suspended in isotonic sucrose plus low concentrations of salt, was measured under steady-state conditions. The relationship between the efflux and the log of the salt concentration can be fitted by two straight lines with a sharp inflection point, the steeper slope occurring at concentrations below 0.2 mM NaCl. The determining factor in the rate of efflux is the ionic strength rather than the specific monovalent cations or anions and the effects are completely reversible. With an increase in temperature, the effects of reduced ionic strength are more pronounced and the inflection point is shifted toward higher salt concentrations. An increase in pH leads to an increased efflux at a given ionic strength, but the size of the pH effect is small at low ionic strength. At a given pH, the data can be fitted by a simplified form of the Goldman equation suggesting that with reduction in ionic strength, the permeability remains constant until the inflection point is reached. At that ionic strength, a sharp reversible transition to a new permeability state occurs. The permeability increases with an increase in the external but not the internal pH.  相似文献   

6.
Despite recent improvement in cellulase enzymes properties, the high cost associated with the hydrolysis step remains a major impediment to the commercialization of full-scale lignocellulose-to-ethanol bioconversion process. As part of a research effort to develop a commercial process for bioconversion of softwood residues, we have examined the potential for recycling enzymes during the hydrolysis of mixed softwood substrate pretreated by organosolv process. We have used response surface methodology to determine the optimal temperature, pH, ionic strength, and surfactant (Tween 80) concentration for maximizing the recovery of bound protein and enzyme activity from the residual substrates after hydrolysis. Data analysis showed that the temperature, pH and surfactant concentration were the major factors governing enzyme desorption from residual substrate. The optimized conditions were temperature 44.4 °C, pH 5.3 and 0.5% Tween 80. The optimal conditions significantly increased the hydrolysis yield by 25% after three rounds of hydrolysis. This bound enzyme desorption combining with free enzyme re-adsorption is a potential method to recover cellulase enzymes and reduce the cost of enzymatic hydrolysis.  相似文献   

7.
Bovine serum albumin (BSA) causes tobacco mosaic virus (TMV) to crystallize at pH values where both have negative charges. The amount of albumin required to precipitate the virus varies inversely with ionic strength of added electrolyte. At pH values above 5, the precipitating power is greatest when BSA has the maximum total, positive plus negative, charge. Unlike early stages of the crystallization of TMV in ammonium sulfate-phosphate solutions, which can be reversed by lowering the temperature, the precipitation of TMV by BSA is not readily reversed by changes in temperature. The logarithm of the apparent solubility of TMV in BSA solutions, at constant ionic strength of added electrolyte, decreases linearly with increasing BSA concentration. This result and the correlation of precipitating power with total BSA charge suggest that BSA acts in the manner of a salting-out agent. The effect of BSA on the reversible entropy-driven polymerization of TMV protein (TMVP) depends on BSA concentration, pH, and ionic strength. In general, BSA promotes TMVP polymerization, and this effect increases with increasing BSA concentrations. The effect is larger at pH 6.5 than at pH 6. Even though increasing ionic strength promotes polymerization of TMVP in absence of BSA, the effect of increasing ionic strength from 0.08 to 0.18 at pH 6.5 decreases the polymerization-promoting effect of BSA. Likewise, the presence of BSA decreases the polymerization-promoting effect of ionic strength. The polymerization-promoting effect of BSA can be interpreted in terms of a process akin to salting-out. The mutual suppression of the polymerization-promoting effects of BSA and of electrolytes by each other can be partially explained in terms of salting-in of BSA.  相似文献   

8.
The ATP-induced enhancement of the intrinsic fluorescence of myosin and heavy meromyosin (HMM) that persists during the steady state of hydrolysis has been investigated. To compare the substrate-induced changes in fluorescence with those in the electron spin resonance spectrum of the spin-labeled enzyme, we studied the influence of temperature, pH, and ionic strength, as well as the effect of chemical modification (spin labeling) of the SH-1 sulfhydryl groups. Changing the pH between 6 and 9 does not affect the enhancement of fluorescence of myosin or HMM; changing the ionic strength, which could be studied only with HMM, also has no effect; and decreasing the temperature from 20 to 5 degrees slightly diminishes the enhancement with both myosin and HMM. Chemical modification with N-(1-oxyl-2,2,6,6-tetramethyl-4-piperidinyl) iodoacetamide, which blocks the SH-1 thiol groups, reduces the enhancement of fluorescence, induces a strong dependence on ionic strength and pH, and substantially increases the dependence on temperature. The enhancement with labeled myosin or labeled HMM increases with increasing pH, ionic strength, and temperature, closely paralleling the effects of these parameters on the electron spin resonance spectrum of spin-labeled myosin (SEIDEL, J.C. and GERGELY, J. (1973) Arch. Biochem. Biophys. 158, 853), suggesting that the same molecular change, induced by ATP and associated with formation of the MADP-P1 complex, underlies both the change in fluorescence and the change in ESR spectrum. Those analogues of ATP that produce the maximal enhancement of fluorescence (WERBER, M., SZENT-GYORGYL, A.G., and FASMAN, G. (1972) Biochemistry 11, 2872) also produce the maximal change in the ESR spectra. Both an amino group at position 6 of the substrate and an unmodified triphosphate chain are required for maximal change in either fluorescence or ESR spectra. The smaller enhancement of fluorescence produced by spin labeling the SH-1 groups persists after the nitroxide has been chemically changed to a diamagnetic species. Thus the small enhancement cannot be attributed to paramagnetic quenching of tryptophan fluorescence by the spin label. An initial burst of phosphate liberation accompanies the hydrolysis of ATP, cytidine 5'-triphosphate, uridine 5'-triphosphate, guanosine 5'-tryphosphate, iosine 5'-triphosphate, 2'-deoxyadenosine 5'-tryphosphate, adenosine 5'-tetraphosphate, and tripolyphosphate. The presence or absence of the burst does not correlate with the extent of the spectral change.  相似文献   

9.
The observed equilibrium constant (Kobs) for the reaction of choline acetyltransferase (EC 2.3.1.6) has been determined under physiological conditions. Using sigma and square brackets to indicate total concentrations of all ionic species present: (see article). The value of Kobs has been determined to be 12.3 plus or minus 0.6 at 38 degrees, pH 7.0 and ionic strength 0.25 M. The value at 25 degrees is not significantly different, and the constant has been found to be insensitive to variations in ionic strength (0.03 to 0.375 M), pH (6.5 TO 7.5) OR FREE [Mg-2+] (0 to 5 mM). The Kobs of this reaction reflects the difference between the observed standard free energy change (delta G-oobs) for the hydrolysis of acetylcholine and the delta G-oobs for the hydrolysis of acetyl-CoA. Since the delta G-oobs for the hydrolysis of acetyl-CoA has been previously determined to be minus 8.54 kcal/mol (minus 35.75 kJ/mol under the same physiological conditions, the delta G-oobs for the reaction of acetylcholinesterase (EC 3.1.1.7): (SEE ARTICLE). Can be calculated to be minus 6.99 kcal/mol (minus 29.26 kJ/mol) at pH ionic strength 0.25 M and 38 degrees, taking the standard state of liquid water to have unit activity ([H2O] equals 1). The pKa for acetic acid under the same conditions, has been determined to be 4.60 plus or minus 0.01, allowing the Kobs for the pH-independent reaction (see article). To be calculated to be 3.28 times 10-2 M. Choline and carnitine are chemical analogues. The Kobs for the corresponding reaction of carnitine acetyltransferase (EC 2.3.1.7). (SEE ARTICLE). Under the same physiological conditions of pH (7.0), ionic strength (0.25 M), and temperature (38 degrees) has been determined to be 1.73 plus or minus 0.05, making the delta G-oobs for the hydrolysis of acetylcholine only 1.21 kcal/mol (5.06 kJ) less negative than that for the hydrolysis of acetylcarnitine.  相似文献   

10.
The monocations of a few 5-substituted cytidines have been shown to undergo competitive deamination to the corresponding uridines and hydrolysis to 5-substituted cytosines and D-ribose. The first-order rate constants measured at different temperatures indicate that the proportion of the hydrolysis is considerably increased with the increasing temperature. Electron-withdrawal by a polar substituent at C5 appears to facilitate the hydrolysis to a larger extent that the deamination. The ionic strength has practically no influence on the rate of either reaction.  相似文献   

11.
The kinetics of binding and hydrolysis of ATP by bovine cardiac myosin subfragment 1 has been reinvestigated. More than 90% of the total fluorescence amplitude associated with ATP hydrolysis occurs with an apparent second-order rate constant of 8.1 X 10(5) M-1 S-1 and a limiting rate constant of approximately 140 S-1 (100 mM KCl, 50 mM 1,3-bis-[tris(hydroxymethyl)methylamino]-propane, 10 mM MgCl2, pH 7.0, 20 degrees C); the remaining 10% occurs more slowly (approximately 1 S-1). The observed rate constants are independent of subfragment 1 concentration under pseudo first-order conditions for ATP with respect to protein. The fraction of protein which hydrolyzes ATP rapidly is not a function of the nucleotide or protein concentration and appears to be constant irrespective of ionic strength or temperature within the range studied (50-100 mM KCl, pH 7.0, 15-20 degrees C). These data are compared to that obtained previously using subfragment 1 prepared by a different method which showed ATP-dependent aggregation of two protein species.  相似文献   

12.
In this study, we have investigated the effect of urea on the physicochemical (pH and conductivity) and rheological properties of the chitosan-β-GP system in order to assess the main polymer–polymer interactions at low and high temperature. The pH of the solutions was slightly increased due to the consumption of H+ in solution through the hydrolysis of urea. Furthermore, the addition of urea considerably decreased the conductivity, and therefore the ionic strength of the solutions, and this effect was more important at high temperature. It indicated that urea strongly affects polymer–polymer interactions by weakening hydrogen bonds at low temperature, but in addition can hinder hydrophobic effect at high temperature since the reduction of ionic strength results in less screening of electrostatic repulsion between protonated glucosamine groups. At 15 °C, the addition of urea to chitosan-β-GP solutions decreased their elasticity, shortened relaxation times and simplified the relaxation process due to the disruption of hydrogen bonds. Heat-induced gelation of the chitosan system in the presence of urea showed higher gelation temperature (Tgel) in non-isothermal tests and longer gelation time (tgel) in isothermal conditions. The activation energy for gelation also increased with increasing urea concentration. We concluded that the detrimental effect of urea on the gelation process was mainly related to a decrease in polymer-polymer hydrophobic effect, as shown by the decrease in conductivity.  相似文献   

13.
Kinetic mechanism of myofibril ATPase.   总被引:18,自引:5,他引:13       下载免费PDF全文
The kinetic mechanism of myofibril ATPase was investigated using psoas and mixed back muscle over a range of ionic strengths. Myofibrils were labeled with pyrene iodoacetamide to measure the rate constants for the binding of ATP and formation of the weakly attached state. The velocity of shortening was measured by stopping the contraction at various times by mixing with pH 4.5 buffer. The transient and steady-state rates of ATP hydrolysis were measured by the quench flow method. The results fitted the kinetic scheme [formula: see text] The rate constants (or equilibrium constants for steps 1 and 6) were obtained for the six steps. k5 was calculated from the KM for shortening velocity, K1, and k2. The rate constants were essentially equal for myofibrils and acto-S-1 at low ionic strength. Increasing the ionic strength up to 100 mM in NaCl increased the rate of the hydrolysis step and the size of the phosphate burst and the effective rate of product release became the rate-limiting step. The step calculated from the velocity of shortening, k5, and k2 is 15 nm, based on a model in which step 4 is the force-generating step.  相似文献   

14.
Alcohol dehydrogenase from horse liver is shown to catalyze ester hydrolysis. Nicotinamide coenzymes do not affect the rate of esterolysis. A kinetic approach to study esterase reaction at low substrate to enzyme ratio is described. Kinetic effects of ester structure, temperature, pH, solvent polarity, and ionic strength were investigated. The liver enzyme enhances the rate of esterolysis by lowering activation energy of reaction according to the Uni-Bi kinetic sequence. Two ionizable groups, cysteine and lysine, are tentatively assigned at the esterolytic site of liver alcohol dehydrogenase from pH-rate profiles and chemical modification studies. A plausible mechanism for the esterase reaction proceeds via the acid-assisted nucleophilic catalysis involving the ammonium ion of lysine and the thiolate of cysteine in the acyl-oxygen cleavage.  相似文献   

15.
The combined effects of ionic strength, divalent cations, pH and toxin concentration on the pore-forming activity of Cry1Ac and Cry1Ca were studied using membrane potential measurements in isolated midguts of Manduca sexta and a brush border membrane vesicle osmotic swelling assay. The effects of ionic strength and divalent cations were more pronounced at pH 10.5 than at pH 7.5. At the higher pH, lowering ionic strength in isolated midguts enhanced Cry1Ac activity but decreased considerably that of Cry1Ca. In vesicles, Cry1Ac had a stronger pore-forming ability than Cry1Ca at a relatively low ionic strength. Increasing ionic strength, however, decreased the rate of pore formation of Cry1Ac relative to that of Cry1Ca. The activity of Cry1Ca, which was small at the higher pH, was greatly increased by adding calcium or by increasing ionic strength. EDTA inhibited Cry1Ac activity at pH 10.5, but not at pH 7.5, indicating that trace amounts of divalent cations are necessary for Cry1Ac activity at the higher pH. These results, which clearly demonstrate a strong effect of ionic strength, divalent cations and pH on the pore-forming activity of Cry1Ac and Cry1Ca, stress the importance of electrostatic interactions in the mechanism of pore formation by B. thuringiensis toxins.  相似文献   

16.
Smooth muscle contraction is controlled in part by the state of phosphorylation of myosin. A recently discovered actin and calmodulin-binding protein, named caldesmon, may also be involved in regulation of smooth muscle contraction. Caldesmon cross-links actin filaments and also inhibits actin-activated ATP hydrolysis by myosin, particularly in the presence of tropomyosin. We have studied the effect of caldesmon on the rate of hydrolysis of ATP by skeletal muscle myosin subfragment-1, a system in which phosphorylation of the myosin is not important in regulation. Caldesmon is a very effective inhibitor of ATP hydrolysis giving up to 95% inhibition. At low ionic strength (approximately 20 mM) this effect does not require smooth muscle tropomyosin, whereas at high ionic strength (approximately 120 mM) tropomyosin enhances the inhibitory activity of caldesmon at low caldesmon concentrations. Cross-linking of actin is not essential for inhibition of ATP hydrolysis to occur since at high ionic strength there is very little cross-linking as determined by a low speed sedimentation assay. Under all conditions examined, the decrease in the rate of ATP hydrolysis is accompanied by a decrease in the binding of myosin subfragment-1 to actin. Furthermore, caldesmon weakens the equilibrium binding of myosin subfragment-1 to actin in the presence of pyrophosphate. We conclude that caldesmon has a general weakening effect on the binding of skeletal muscle myosin subfragment-1 to actin and that this weakening in binding may be responsible for inhibition of ATP hydrolysis.  相似文献   

17.
The equilibrium constant of the phosphoglyceromutase reaction was determined over a range of pH (5.4-7.9), in solutions of different ionic strength (0.06-0.3) and in the presence of Mg(2+), at 30 degrees C and at 20 degrees C. The values obtained (8.65-11.65) differ substantially from previously published values. The third acid dissociation constants were redetermined for 2- and 3-phosphoglycerate, and in contrast with previous reports the pK values (7.03 and 6.97 respectively at zero ionic strength) were closely similar. The Mg(2+)-binding constants were measured spectrophotometrically and the values, 286mm(-1) and 255mm(-1) for 2- and 3-phosphoglycerate at pH7 and ionic strength 0.02, were also very similar. From the relative lack of effect of temperature, pH and ionic strength it is concluded that the equilibrium constant differs from unity largely because of entropic factors. At low ionic strength, in the neutral region, the pH-dependence can be attributed to the small difference in the acid dissociation constants, but the difference in dissociation constants does not explain the pH-dependence in the acid region or at high ionic strength. Within physiological ranges of pH, Mg(2+) concentration and ionic strength there will be little variation in equilibrium constant.  相似文献   

18.
Renaturation of denatured, covalently closed circular DNA   总被引:2,自引:0,他引:2  
The rate of renaturation of denatured, covalently closed, circular DNA (form Id DNA) of the phi X174 replicative form has been investigated as a function of pH, temperature, and ionic strength. The rate at a constant temperature is a sharply peaked function of pH in the range of pH 9 to 12. The position on the pH scale of the maximum rate decreases as the temperature is increased and as the ionic strength is increased. The kinetic course of renaturation is pseudo-first order: it is independent of DNA concentration, but falls off in rate from a first order relationship as the reaction proceeds. The rate of renaturation depends critically on the temperature at which the denaturation is carried out. Form Id, prepared at an alkaline pH at 0 degrees C, renatures from 5 to more than 100 times more rapidly than that similarly prepared at 50 degrees C. Both the heterogeneity in rate and the effect of the temperature of denaturation depend, in part, on the degree of supercoiling of the form I DNA from which the form Id is prepared. However, it is concluded that a much larger contribution to both arises from a configurational heterogeneity introduced in the denaturation reaction. The renaturation rate was determined by neutralization of the alkaline reaction and analytical ultracentrifugal analysis of the amounts of forms I and Id. The nature of the proximate renatured species at the temperature and alkaline pH of renaturation was investigated by spectrophotometric titration and analytical ultracentrifugation. It is concluded that the proximate species are the same as the intermediate species defined by an alkaline sedimentation titration of the kind first done by Vinograd et al. ((1965) Proc. Natl. Acad. Sci. U. S. A. 53, 1104-1111). Observations are included on the buoyant density of form Id and on depurination of DNA at alkaline pH values and high temperatures.  相似文献   

19.
Talin is an adhesion plaque protein believed important in linking actin filaments to the plasma membrane. The nature of a direct talin-actin interaction, however, is complex and has remained unclear. We have systematically characterized the effects of pH, ionic strength, temperature, and protein molar ratio on the interaction between highly purified talin and actin. The ability of talin to increase viscosity of F-actin at 25 degrees C and low ionic strength increased with decreasing pH from 7.3 to 6.4 and increasing molar ratio of talin to actin. At pH 6.4 and low ionic strength, talin could extensively crosslink actin filaments into ordered bundles as shown by negative staining and could cosediment with F-actin at molar ratios as high as one talin to two actin monomers. Talin crosslinked prepolymerized actin filaments to a similar extent as actin filaments polymerized in its presence. The 190-kDa calpain-generated proteolytic fragment of talin bound poorly to actin under conditions favorable for intact talin, but was able to crosslink actin filaments at a lower pH. Increasing the ionic strength within a relatively narrow range significantly decreased ability of talin to bind to actin, regardless of pH. The effects of pH and ionic strength on the talin-actin interaction were rapid and reversible. Low-shear-viscosity studies revealed a strong temperature dependence in the talin-actin interaction with significant crosslinking activity at physiological-like ionic conditions and temperature (37 degrees C). Our results consistently demonstrated that talin crosslinks actin filaments and that this direct interaction is highly sensitive to, and dependent upon, ionic conditions and temperature.  相似文献   

20.
The kinetics of the Cu(II) catalysed hydrolysis of thiamine-pyrophosphate (TPP) has been studied in aqueous solution at 56,64 and 78° over a pH fange of 3.0 to 7.0 at a constant ionic strength of 0.10 M (KNO3). The pH rate profiles were analysed and the overall rate constants resolved into individual specific rate constants relating to various Cu:TPP chelate species in solution. Activation parameters ΔH2, ΔS° and ΔG2 for the specific rate constants of various chelate species of TPP are reported. The possible mechanism of the Cu(II) promoted hydrolysis of TPP is discussed. The structure-reactivity relationship is also discussed.  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