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 mechanisms involving inactivation of bacteriophage MS2 in aerosols and the effect of protective substances in the spray-medium were studied after spraying from various NaCl solutions. Results with aerosols generated from the salt solutions showed that with higher salt concentration in the spray-medium higher concentrations of protective substances were needed to protect phage MS2 against aerosol inactivation. Phenylalanine, which has a protective action at low concentration, produced less protection in aerosol droplets that were supersaturated solutions of this substance or in which crystals of phenylalanine can be expected to form. Our results suggested that protection by peptone and phenylalanine was related to the concentration in the aerosol droplet after evaporation to equilibrium, whereas protection by the surface active agent OED (a commercial mixture of oxyethylene docosylether and oxyethelene octadecylether) was related to the concentration at which a monolayer is formed around the aerosol particle. Inactivation of phage MS2 was maximal in the aerosol particle in fluid phase and became less at lower relative humidity where aerosol particles are expected to be in the solid state. It is suggested that inactivation of bacteriophage MS2 in aerosols could be explained by surface inactivation at the air-water interface.  相似文献   

2.
1. If dilute solutions of purified trypsin of low salt concentration at pH from 1 to 7 are heated to 100°C. for 1 to 5 minutes and then cooled to 20°C. there is no loss of activity or formation of denatured protein. If the hot trypsin solution is added directly to cold salt solution, on the other hand, all the protein precipitates and the supernatant solution is inactive. 2. The per cent of the total protein and activity present in the soluble form decreases from 100 per cent to zero as the temperature is raised from 20°C. to 60°C. and increases again from zero to 100 per cent as the solution is cooled from 60°C. to 20°C. The per cent of the total protein present in the soluble (native) form at any one temperature is nearly the same whether the temperature is reached from above or below. 3. If trypsin solutions at pH 7 are heated for increasing lengths of time at various temperatures and analyzed for total activity and total protein nitrogen after cooling, and for soluble activity and soluble (native) protein nitrogen, it is found that the soluble activity and soluble protein nitrogen decrease more and more rapidly as the temperature is raised, in agreement with the usual effects of temperature on the denaturation of protein. The total protein and total activity, on the other hand, decrease more and more rapidly up to about 70°C. but as the temperature is raised above this there is less rapid change in the total protein or total activity and at 92°C. the solutions are much more stable than at 42°C. 4. Casein and peptone are not digested by trypsin at 100°C. but when this digestion mixture is cooled to 35°C. rapid digestion occurs. A solution of trypsin at 100°C. added to peptone solution at zero degree digests the peptone much less rapidly than it does if the trypsin solution is allowed to cool slowly before adding it to the peptone solution. 5. The precipitate of insoluble protein obtained from adding hot trypsin solutions to cold salt solutions contains the S-S groups in free form as is usual for denatured protein. 6. The results show that there is an equilibrium between native and denatured trypsin protein the extent of which is determined by the temperature. Above 60°C. the protein is in the denatured and inactive form and below 20°C. it is in the native and active form. The equilibrium is attained rapidly. The results also show that the formation of denatured protein is proportional to the loss in activity and that the re-formation of native protein is proportional to the recovery of activity of the enzyme. This is strong evidence for the conclusion that the proteolytic activity of the preparation is a property of the native protein molecule.  相似文献   

3.
Leupeptin (acyl peptidyl-L-argininal) is a potent inhibitor of trypsin and related proteases. We analyzed the association of leupeptim with bovine trypsin kinetically, assuming that it proceeds by a pathway which involves two steps: E + I in equilibrium K1 Complex I k-2 in equilibrium k+2 Complex II. The observed dissociation constant (K1) for the first step was 1.24 X 10(-3) M (at pH 8.2 15 degrees C) and the two first-order rate constants (k+2 and k-2) were 166 s-1 and 1.75 X 10(-3.s-1, respectively (at pH 8.2, 15 degrees C). The dissociation constant (Kd) for the whole process was calculated from these parameters to be 1.34 X 10(-8) M. This value is compatible with that determined directly by an independent static method (2.36 X 10(-8) M). We also measured Kd for the leupeptine complex of anhydrotrypsin, a trypsin derivative in which the active-site hydroxyl group is missing. The observed value was about 5 orders of magnitude larger than Kd and was rather similar to K1 in native trypsin. A elupeptin isomer which contains a D-argininal residue did not show strong affinity towards trypsin. These findings suggest that complex II consists of a covalent hemiacetal adduct formed between the serine hydroxyl group in the enzyme active site and the aldehyde group in the inhibitor. The pH dependencies of the dissociation constant and other parameters show that deprotonation of the charge-relay sustem in the active site is important for the formation and stabilization of complex II.  相似文献   

4.
The proteolytic activity of dilute solutions of clystalline trypsin is destroyed by x-rays, the amount of inactivation being an exponential function of the radiation dose. The reaction yield increases steadily with increasing concentration of trypsin, varying, as the concentration of enzyme is increased from 1 to 300 microM, from 0.068 to 0.958 micromole of trypsin per liter inactivated per 1000 r with 0.005 N hydrochloric acid as the solvent, from 0.273 to 0.866 with 0.005 N sulfuric acid as the solvent, and from 0.343 to 0.844 with 0.005 N nitric acid as the solvent. When the reaction yields are plotted as a function of the initial concentration of trypsin, they fall on a curve given by the expression Y alpha X(K), in which Y is the reaction yield, X is the concentration of trypsin, and K is a constant equal to 0.46, 0.20, and 0.16, respectively, with 0.005 N hydrochloric, sulfuric, and nitric acids as solvents. The differences between the reaction yields found with chloride and sulfate ions in I to 10 microM trypsin solutions are significant only in the pH range from 2 to 4. The amount of inactivation obtained with a given dose of x-rays depends on the pH of the solution being irradiated and the nature of the solvent. The reaction yield-pH curve is a symmetrical one, with minimum yields at about pH 7. Buffers such as acetate, citrate, borate and barbiturate, and other organic molecules such as ethanol and glucose, in concentrations as low as 20 microM, inhibit the inactivation of trypsin by x-radiation. Sigmoid inactivation-dose curves instead of exponential ones are obtained in the presence of ethanol. The reaction yields for the inactivation of trypsin solutions by x-rays are approximately 1.5 times greater when the irradiation is done at 26 degrees C. than when it is done at 5 degrees C., when 0.005 N hydrochloric acid is the solvent. The dependence on temperature is less when 0.005 N sulfuric acid is used, and is negligible with 0.005 N nitric acid. The difficulties involved in interpreting radiation effects in aqueous systems, and in comparing the results obtained under different experimental conditions, are discussed.  相似文献   

5.
1. The rate of inactivation of crystalline trypsin solutions and the nature of the products formed during the inactivation at various pH at temperatures below 37°C. have been studied. 2. The inactivation may be reversible or irreversible. Reversible inactivation is accompanied by the formation of reversibly denatured protein. This denatured protein exists in equilibrium with the native active protein and the equilibrium is shifted towards the denatured form by raising the temperature or by increasing the alkalinity. The decrease in the fraction of active enzyme present (due to the formation of this reversibly denatured protein) as the pH is increased from 8.0 to 12.0 accounts for the decrease in the rate of digestion of proteins by trypsin in this range of pH. 3. The loss of activity at high temperatures or in alkaline solutions, just described, is very rapid and is completely reversible for a short time only. If the solutions are allowed to stand the loss in activity becomes gradually irreversible and is accompanied by the appearance of various reaction products the nature of which depends upon the temperature and pH of the solution. 4. On the acid side of pH 2.0 the trypsin protein is changed to an inactive form which is irreversibly denatured by heat. The course of the reaction in this range is monomolecular and its velocity increases as the acidity increases. 5. From pH 2.0 to 9.0 trypsin protein is slowly hydrolyzed. The course of the inactivation in this range of pH is bimolecular and its velocity increases as the alkalinity increases to pH 10.0 and then decreases. As a result of these two reactions there is a point of maximum stability at about pH 2.3. 6. On the alkaline side of pH 13.0 the reaction is similar to that in strong acid solution and consists in the formation of inactive protein. The course of the reaction is monomolecular and the velocity increases with increasing alkalinity. From pH 9.0 to 12.0 some hydrolysis takes place and some inactive protein is formed and the course of the reaction is represented by the sum of a bi- and monomolecular reaction. The rate of hydrolysis decreases as the solution becomes more alkaline than pH 10.0 while the rate of formation of inactive protein increases so that there is a second point at about pH 13.0 at which the rate of inactivation is a minimum. In general the decrease in activity under all these conditions is proportional to the decrease in the concentration of the trypsin protein. Equations have been derived which agree quantitatively with the various inactivation experiments.  相似文献   

6.
Human tryptase-beta (HTbeta) is a serine protease that is isolated as a tetramer of four identical, catalytically active subunits (HTbeta-AT). Tetramer activity is not affected by protein-based physiological inhibitors but instead may be regulated by an autoinactivation process we have called spontaneous inactivation. Unless stabilized by heparin or high salt, the active tetramer converts to an inactive state consisting of an inactive-destabilized tetramer that reversibly dissociates to inactive monomers upon dilution. We refer to this mixture of inactive species as siHTbeta and show in this study that previous reports of monomeric catalytic forms are derived from this mixture. siHTbeta itself did not hydrolyze model substrates but unlike the tetramer did react slowly with the serpin alpha2-antiplasmin (alpha2-AP), suggesting a highly limited catalytic potential. In the presence of heparin (or other highly charged polysaccharides), we demonstrate that siHTbeta formed a well-defined complex with the heparin (siHTbeta-HC) that reacted 70-fold faster with alpha2-AP than siHTbeta and also hydrolyzed model substrates and fibrinogen. Formation of siHTbeta-HC was limited to dilute subunit solutions since high subunit concentrations resulted in the reformation of the active tetramer. By compensating for changes in the strength of heparin binding, siHTbeta-HC could be formed over the pH range of 6.0-8.5. The activity dependence on pH was bell-shaped with highest activity between pH 6.8 and pH 7.5. In contrast, HTbeta-AT activity showed no dependence upon heparin, increased over the pH range of 6.0-8.5, and was much higher than that of siHTbeta-HC especially above pH 6.8. HTbeta-AT incubated with excess heparin of different size (3-15 kDa) was functionally stable at 25 degrees C but lost activity regardless of heparin size at 37 degrees C above pH 6.8. The change in stability, which is likely due to weakened heparin binding, did not result in the formation of a stable catalytic monomer. These results confirm that siHTbeta is for the most part an inactive species and that any active monomer is a consequence of heparin binding to siHTbeta under dilute conditions where unfavorable thermodynamics and/or kinetics restrict formation of active tetramer. Heparin binding under these conditions drives a limited reorganization of the active site to a conformation that is catalytic but not the equivalent of a subunit within the active tetramer.  相似文献   

7.
The stability of bacterial viruses in solutions of salts   总被引:2,自引:2,他引:0  
1. The seven bacterial viruses of the T group, active against E. coli, are much more rapidly inactivated by heat when suspended in 0.1 N solutions of sodium salts than when suspended in broth. 2. The kinetics of this inactivation whether in salt solutions or in broth are those of a first order reaction. 3. The rate of inactivation of phage T5 in 0.1 N NaCl at 37 degrees C. can be greatly decreased by the addition of 10(-8)M concentrations of such divalent cations as Ca, Mg, Ba, Sr, Mn, Co, Ni, Zn, Cd, and Cu. 4. An increase in the cation concentration in the suspending medium results in an increase in the stability of phage T5 to the inactivating effects of temperature. 5. The hypothesis is proposed that the increase in stability of phage T5 in the presence of various cations is the result of complex formation between the phage and the metal ion.  相似文献   

8.
The activity of dilute solutions of crystalline trypsin is destroyed by x-rays. The inactivation is an exponential function of the radiation dose. The reaction yield of inactivation is independent of the intensity at which the radiation is delivered or the quality of the x-rays. The reaction yield increases with increasing concentration of trypsin, varying from 0.06 to 0.7 micromoles per liter per 1000 r for trypsin solutions ranging from 1 x 10(-7) to 2 x 10(-4)M.  相似文献   

9.
Adenosine-5'-phosphosulfate (APS) kinase from Penicillium chrysogenum, loses catalytic activity at temperatures greater than approximately 40 degrees C. When the heat-inactivated enzyme is cooled to 30 degrees C or lower, activity is regained in a time-dependent process. At an intermediary temperature (e.g. 36 degrees C) an equilibrium between active and inactive forms can be demonstrated. APS kinase from P. chrysogenum is a dimer (Mr = 57,000-60,000) composed of two apparently identical subunits. Three lines of evidence suggest that the reversible inactivation is a result of subunit dissociation and reassociation. (a) Inactivation is a first-order process. The half-time for inactivation at a given temperature is independent of the original enzyme concentration. Reactivation follows second-order kinetics. The half-time for reactivation is inversely proportional to the original enzyme concentration. (b) The equilibrium active/inactive ratio at 36 degrees C increases as the total initial enzyme concentration is increased. However, Keq,app at 5 mM MgATP and 36 degrees C calculated as [inactive sites]2/0.5 [active sites] is near-constant at about 1.7 X 10(-8) M over a 10-fold concentration range of enzyme. (c) At 46 degrees C, the inactive P. chrysogenum enzyme (assayed after reactivation) elutes from a calibrated gel filtration column at a position corresponding to Mr = 33,000. Substrates and products of the APS kinase reaction had no detectable effect on the rate of inactivation. However, MgATP and MgADP markedly stimulated the reactivation process (kapp = 3 X 10(5) M-1 X s-1 at 30 degrees C and 10 mM MgATP). The kapp for reactivation was a nearly linear function of MgATP up to about 20 mM suggesting that the monomer has a very low affinity for the nucleotide compared to that of the native dimer. Keq,app at 36 degrees C increases as the MgATP concentration is increased. The inactivation rate constant increased as the pH was decreased but no pK alpha could be determined. The reactivation rate constant increased as the pH was increased. An apparent pK alpha of 6.4 was estimated.  相似文献   

10.
Kinetics of trypsin association with trypsin inhibitor from colostrum (IC) was studied. The association rate constant is 3-10-5 M- minus 1 sec- minus 1 at pH 7,8, 25 degrees C. The rate constant for the complex dissociation was determined from the kinetics of the IC displacement from the complex with trypsin by a specific substrate and was found to be 5-10- minus 6 sec- minus 1 (pH 7,8; 25 degrees C). The equilibrium constant (Ki) was measured in a special experiment and was equal to 4-10- minus 12 M (p H 7,8; 25 degrees C). The similarity of this reaction and the association of trypsin with other protein inhibitors was discussed.  相似文献   

11.
We have previously described [H, Kühn et al. (1981) Proc. Natl Acad. Sci. USA, 78, 6873-6877] a light-induced scattering change ('binding signal') associated with a stoichiometric binding between photoexcited rhodopsin and a peripheral membrane protein, the GTP-binding protein, in bovine rod outer segment suspensions. We have attempted here to identify the rhodopsin intermediate R* which is responsible for this interaction, by studying its dependence on pH, temperature and ionic strength. The results strongly suggest that the active state is metarhodopsin II (M II). 1. The initial phase of the binding signal is slightly slower than the formation of metarhodopsin II (2-37 degrees C, pH 5.5-9). 2. The kinetics of the decay of the active rhodopsin state are similar to those of the metarhodopsin II leads to metarhodopsin III transition (37 degrees C, pH 7.3). 3. All conditions which lead to light-induced binding of the GTP-binding protein to R* also lead to the formation of M II. At 2 degrees C, pH 8.3, in particular where no M II is formed in the absence of GTP-binding protein, binding signals and light-induced attachment of the GTP-binding protein to the membrane are still observed. Consistently, addition of GTP-binding protein to a suspension of extracted membranes bleached at 2 degrees C (pH 8.3) shifts the metarhodopsin I in equilibrium metarhodopsin II equilibrium towards metarhodopsin II. The shift is reversed by GTP, which dissociates the rhodopsin--GTP-binding protein complex. 4. At low ionic strength, where the GTP-binding protein is soluble in the dark (instead of being associated to the membrane as in the above experiments) M II still induces the binding whereas M I does not, indicating a much lower affinity of the GTP-binding protein for MI.  相似文献   

12.
General aspects of the mechanism of antithrombin action were elucidated by a comparison of the inactivation of trypsin by antithrombin with the inactivation of coagulation proteinases by the inhibitor. Bovine antithrombin and bovine trypsin were shown to form an inactive equimolar complex. A non-complexed, proteolytically modified form of antithrombin, electrophoretically identical with that formed in the reaction with coagulation proteinases, was also produced in the reaction with trypsin. In the absence of heparin, the inactivation of trypsin by antithrombin was 20 times faster than the inactivation of thrombin; the second-order rate constant was 1.5 x 10(5)m(-1).s(-1) at 25 degrees C and pH 7.4. However, the inhibition of thrombin was accelerated about 30 times more efficiently by small amounts of heparin than was trypsin inhibition. Dissociation of the antithrombin-trypsin complex at pH 7.4 followed first-order kinetics with a half-life for the complex of about 80h at 25 degrees C. The complex was rapidly and quantitatively dissociated at pH 11, resulting in the liberation of a modified two-chain form of the inhibitor, cleaved at the same Arg-Ser bond as in modified antithrombin released from complexes with thrombin, Factor Xa and Factor IXa. This supports the previous proposal that this bond is the active-site bond of antithrombin. Antisera specific for thrombin-modified antithrombin reacted with purified antithrombin-trypsin complex, indicating that the inhibitor was present in the complex in a form immunologically identical with thrombin-modified antithrombin. The results thus suggest a common mechanism, but different kinetics, for the inhibition of trypsin and coagulation proteinases by antithrombin.  相似文献   

13.
RNA binding properties of the coat protein from bacteriophage GA.   总被引:2,自引:0,他引:2       下载免费PDF全文
The coat protein of bacteriophage GA, a group II RNA phage, binds to a small RNA hairpin corresponding to its replicase operator. Binding is specific, with a Ka of 71 microM -1. This interaction differs kinetically from the analogous coat protein-RNA hairpin interactions of other RNA phage and also deviates somewhat in its pH and salt dependence. Despite 46 of 129 amino acid differences between the GA and group I phage R17 coat proteins, the binding sites are fairly similar. The essential features of the GA coat protein binding site are a based-paired stem with an unpaired purine and a four nucleotide loop having an A at position -4 and a purine at -7. Unlike the group I phage proteins, the GA coat protein does not distinguish between two alternate positions for the unpaired purine and does not show high specificity for a pyrimidine at position -5 of the loop.  相似文献   

14.
Previous X-ray studies (2.8-A resolution) on the crystals of tobacco mosaic virus protein (TMVP) grown from solutions containing high salt have characterized the structure of the protein aggregate as a bilayered cylindrical disk formed by 34 identical subunits [Bloomer, A.C., Champness, J.N., Bricogne, G., Staden, R., & Klug, A. (1978) Nature (London) 276, 362-368]. Under low-salt conditions, 20S aggregates are in equilibrium with 4S species and involved in the efficient nucleation of TMV assembly in vitro [Butler, P.J.G. (1984) J. Gen. Virol. 65, 253-279]. We have investigated by sedimentation velocity and near-UV circular dichroism (CD) measurements the structure of 20S aggregates in low salt (I = 0.1 potassium phosphate at pH 7.0 and 20 degrees C) and the aggregates in high salt [0.2 M (NH4)2SO4 in I = 0.1 tris(hydroxymethyl)aminomethane hydrochloride at pH 8.0 and 20 degrees C, close to the conditions under which TMVP crystallizes as disk aggregates]. At high salt, we observe structures (presumably stacks of disks) having s20,w values around 40, 45, and 50 S, but not the 20S species present in low-salt buffers. The near-UV CD spectrum of 20S aggregates has been obtained for the first time, using computer techniques, from the spectra of the 4S-20S equilibrium mixture and the 4S species. This spectrum of 20S aggregates differs dramatically from that of the stacks of disks examined at both high and low salt (into which the stacks can be returned by dialysis), indicating that the difference is not a solvent effect.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
1. The binding of sodium n-dodecyl sulphate to trypsin and reduced trypsin has been measured by equilibrium dialysis at pH 3.5 and 5.5. 2. At pH 3.5 trypsin specifically binds surfactant at low concentration, at higher concentrations co-operative binding occurs. 3. Reduction of trypsin destroys the specific binding sites at pH 3.5. 4. At pH 5.5 both trypsin and reduced trypsin show only co-operative binding. 5. The interaction of sodium n-dodecyl sulphate with trypsin, reduced, inhibited, and thermally denatured trypsins has been studied by microcalorimetry at 25 degrees C. 6. The microcalorimetric measurements have been used to estimate enthalpy changes (deltaHd) on unfolding of trypsin; deltaHd = 82 +/- 5 kJ-mol-1 at pH 3.5 and 128 +/- 5 kJ-mol-1 at pH 5.5. 7. The unfolding of trypsin follows a different thermochemical pathway to that of reduced trypsin.  相似文献   

16.
Catalytic amounts of bovine beta-trypsin, bovine alpha-chymotrypsin and porcine plasmin establish a true thermodynamic equilibrium between virgin (I) (reactive site Lys15-Ala16 peptide bond intact) and modified (I) (this bond hydrolyzed) bovine trypsin/kallikrein inhibitor (Kunitz). The very slow reaction rates for attaining equilibrium are pH-dependent and differ for different enzymes. Optimal rates are for beta-trypsin at pH 3.75, for alpha-chymotrypsin at pH 5.5, and for plasmin at pH 5.0. Under conditions of optimum pH the equilibrium is reached with the highest rate by plasmin. In 10(-5)M inhibitor solutions the equilibrium concentrations of virgin and modified inhibitor are established by plasmin after almost 300 days starting from either pure virgin or pure modified inhibitor. Thus, the hydrolysis constant KHyd = [I]/[I] is determined to be 0.33 at pH 5.0. In spite of many unsuccessful attempts, this demonstrates that the reactive site peptide bond Lys15-Ala16 in the bovine trypsin inhibitor (Kunitz) can be hydrolyzed by catalytic amounts of endopeptidase. It further confirms that the hydrolyzed Lys15-Ala16 peptide bond in modified inhibitor is subject to thermodynamic control resynthesis.  相似文献   

17.
The adenosinetriphosphatase (ATPase) (EC 3.6.1.3) activity in Azotobacter vinelandii concentrates in the membranous R3 fraction that is directly associated with Azotobacter electron transport function. Sonically disrupted Azotobacter cells were examined for distribution of ATPase activity and the highest specific activity (and activity units) was consistently found in the particulate R3 membranous fraction which sediments on ultracentrifugation at 144 000 X g for 2 h. When the sonication time interval was increased, the membrane-bound ATPase activity could neither be solubilized nor released into the supernatant fraction. Optimal ATPase activty occurred at pH 8.0; Mg2+ ion when added to the assay was stimulatory. Maximal activity always occurred when the Mg2+:ATP stoichiometry was 1:1 on a molar ratio at the 5 mM concentration level. Sodium and potassium ions had no stimulatory effect. The reaction kinetics were linear for the time intervals studied (0-60 min). The membrane-bound ATPase in the R3 fraction was stimulated 12-fold by treatment wiTH TRypsin, and fractionation studies showed that trypsin treatment did not solubilize ATPase activity off the membranous R3 electron transport fraction. The ATPase was not cold labile and the temperature during the preparation of the R3 fraction had no effect on activity; overnight refrigeration at 4 degrees C, however, resulted in a 25% loss of activity as compared with a 14% loss when the R3 fraction was stored overnight at 25 degrees C. A marked inactivation (although variable, usually about 60%) did occur by overnight freezing (-20 degrees C), and subsequent sonication failed to restore ATPase activity. This indicates that membrane reaggregation (by freezing) was not responsible for ATPase inactivation. The addition of azide, ouabain, 2,4-dinitrophenol, or oligomycin to the assay system resulted in neither inhibition nor stimulation of the ATPase activity. The property of trypsin activation and that ATPase activity is highest in the R3 electron transport fraction suggests that its probable functional role is in coupling of electron transport to oxidative phosphorylation.  相似文献   

18.
Stability of minute virus of mice against temperature and sodium hydroxide.   总被引:1,自引:0,他引:1  
Treatment with steam and/or dilute NaOH are commonly used techniques to disinfect manufacturing vessels and tools in the pharmaceutical industry. The aim of this procedure is sanitisation and inactivation of microbiological and viral contaminants. Here we describe the inactivation of the mouse parvovirus Minute Virus of Mice (MVM) under these conditions. Parvoviruses are known to be resistant to physico-chemical treatment and one representative of this family, the human parvovirus B19, is a potential contaminant of blood plasma. We show inactivation kinetics for MVM treated with wet-heat (70, 80, 90 degrees C) and with 0.01-1 M NaOH solutions (pH >/=11.9). Robust inactivation was only achieved at 90 degrees C for at least 10 min and in NaOH solutions of pH >/=12.8 (0.1 M NaOH). It was observed, that aggregation of viruses might protect viral particles from inactivation by NaOH. Therefore, appropriate sample preparation of spiking material is important for accurate simulation of the naturally occurring situation. The observed stability at pH 11.8 exceeds the previously reported upper limit of pH 9. Inactivation was due to disintegration of the viral capsid as assessed by accessibility of viral DNA for endonucleases.  相似文献   

19.
The rate of mitochondrial ATPase inactivation by the naturally occurring inhibitor protein in the presence of saturating ATP and Mg2+ at pH 8.0 depends hyperbolically on the amount of inhibitor added; the upper limit of an apparent first-order constant for the inactivation process is 1.0(-1) at 25 degrees C. A dramatic difference in the inactivation rate is observed when the protein inhibitor is added to the same assay system from either acidic (pH 4.8) or alkaline (pH 8.2) solutions. The slow reversible transition of the inhibitor from its rapidly reacting 'acidic' form to the slow reacting 'alkaline' form occurs when the solution of the protein inhibitor is subjected to a pH-jump from 4.8 to 8.2 (t1/2 approximately 30s at 25 degrees C). The pH-profile of the inhibitor active/inactive equilibrium suggests that a group with pKa approximately 6.5 is involved in the transition. Treatment of the inhibitor protein with a histidine-specific reagent (e.g. diethyl pyrocarbonate) abolishes its inactivating effect on the ATPase activity. It is concluded that the protonation/deprotonation of the inhibitor protein followed by its slow conformational changes is the rate-limiting step in the inhibitor-ATP synthetase interaction.  相似文献   

20.
The kinetics of formation and of dissociation of open complexes (RPo) between Escherichia coli RNA polymerase (R) and the lambda PR promoter (P) have been studied as a function of temperature in the physiological range using the nitrocellulose filter binding assay. The kinetic data provide further evidence for the mechanism R + P in equilibrium I1 in equilibrium I2 in equilibrium RPo, where I1 and I2 are kinetically distinguishable intermediate complexes at this promoter which do not accumulate under the reaction conditions investigated. The overall second-order association rate constant (ka) increases dramatically with increasing temperature, yielding a temperature-dependent activation energy in the range 20 kcal (near 37 degrees C) to 40 kcal (near 13 degrees C) (1 kcal = 4.184 kJ). Both isomerization steps (I1----I2 and I2----RPo) appear to be highly temperature dependent. Except at low temperatures (less than 13 degrees C) the step I1----I2, which we attribute to a conformational change in the polymerase with a large negative delta Cp degrees value, is rate-limiting at the reactant concentrations investigated and hence makes the dominant contribution to the apparent activation energy of the pseudo first-order association reaction. The subsequent step I2----RPo, which we attribute to DNA melting, has a higher activation energy (in excess of 100 kcal) but only becomes rate-limiting at low temperature (less than 13 degrees C). The initial binding step R + P in equilibrium I1 appears to be in equilibrium on the time-scale of the isomerization reactions under all conditions investigated; the equilibrium constant for this step is not a strong function of temperature and is approximately 10(7) M-1 under the standard ionic conditions of the assay (40 mM-Tris . HCl (pH 8.0), 10 mM-MgCl2, 0.12 M-KC1). The activation energy of the dissociation reaction becomes increasingly negative at low temperatures, ranging from approximately -9 kcal near 37 degrees C to -30 kcal near 13 degrees C. Thermodynamic (van't Hoff) enthalpies delta H degrees of open complex formation consequently are large and temperature-dependent, increasing from approximately 29 to 70 kcal as the temperature is reduced from 37 to 13 degrees C. The corresponding delta Cp degrees value is approximately -2.4 kcal/deg. We propose that this large negative delta Cp degrees value arises primarily from the burial of hydrophobic surface in the conformational change (I1 in equilibrium I2) in RNA polymerase in the key second step of the mechanism.(ABSTRACT TRUNCATED AT 400 WORDS)  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