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 15 毫秒
1.
Bile salt stimulated human milk lipase (E.C.3.1) has been used to catalyze the hydrolysis of 4-nitrophenylalkanoates (alkyl C-chain length, n = 2, 3, 4, 6, 10, 12 and 16) in a detergentless microemulsion medium of n-hexane/iso-propanol/water (71.1:27.7:1.2 vol%) containing 2 mM taurocholate at 37°C. The rate of hydrolysis of the esters with n = 2, 3, or 4 was nearly equal but the rate then fell rapidly with increasing alkyl-chain length. No activity was observed with the palmitate ester. Three dimensional surfaces were built up to represent both the catalytic activity and the rate constant of inactivation for the enzyme catalyzed hydrolysis of 4-nitrophenylpropionate in a range of compositions of the ternary system n-hexane/wo-propanol/water at 37 °C. Maximum activation and maximum inactivation were observed in the ternary region of the phase diagram. In the stable; transparent micro-emulsion region another smaller maximum in activity was observed and, at this same solvent composition, minimum inactivation occurred. Values of Km lay in the range 1.37–8.98 mM while Vmax varied from 0.25–7.59 umol.min.mg?1 and the rate constant of inactivation kin ranged from (0.01–1.18). 10?3 s?1 over the three-dimensional surfaces. The most important result is that the enzyme retains its activity in microemulsion media containing less than two volume percent water content.  相似文献   

2.
Catalytic activity and stability of cholesterol oxidase dissolved in ternary systems composed of n-hexane, isopropanol, and water were studied. The dependence of catalytic activity on the composition of the system revealed two maxima, in contrast to the behaviour of previously studied enzymes where a single maximum has been observed. The stability profile of cholesterol oxidase showed a single sharp maximum coinciding with the microemulsion region of the phase diagram. Both catalytic activity and the first-order inactivation rate constant of cholesterol oxidase dissolved in n-hexane/isopropanol/water ternary systems were found to decrease with decreasing temperature. This decrease was more rapid for the inactivation rate constant than for catalytic activity, the activation energies being 200 and 60 kJ.mol-1, respectively. Preparative conversion of cholesterol to cholestenone catalyzed by cholesterol oxidase in n-hexane/isopropanol/water ternary systems was carried out with 100% yield. Decreased temperature and the presence of catalase were required to achieve high degrees of cholesterol conversion. A simple procedure suitable for rapid separation of the reaction product and recovery of the enzyme was developed.  相似文献   

3.
A comparative study of thermal denaturation and inactivation of aspartate aminotransferase from pig heart mitochondria (mAAT) has been carried out (10 mM Na phosphate buffer, pH 7.5). Analysis of the data on differential scanning calorimetry shows that thermal denaturation of mAAT follows the kinetics of irreversible reaction of the first order. The kinetics of thermal inactivation of mAAT follows the exponential law. It has been shown that the inactivation rate constant (kin) is higher than the denaturation rate constant (kden). The kin/kden ratio decreases from 28.8 ± 0.1 to 1.30 ± 0.09 as the temperature increases from 57.5 to 77 °C. The kinetic model explaining the discrepancy between the inactivation and denaturation rates has been proposed. The size of the protein aggregates formed at heating of mAAT at a constant rate (1 °C min− 1) has been characterized by dynamic light scattering.  相似文献   

4.
Brevibacterium imperialis CBS 489-74 was grown in broths prepared with yeast and malt extract, bacteriological peptone and 2% glucose or differently modified with the addition of Na-phosphate buffer, FeSO4, MgSO4 and CoCl2. The peak production of nitrile hydratase (NHase) did not change significantly. At the stationary growth phase, the units per milliliter of broth (60 units ml−1) were more important than those at the exponential growth phase.

The NHase operational stability of whole resting cells was monitored following the bioconversion of acrylonitrile to acrylamide in continuous and stirred UF-membrane reactors. The rate of inactivation was independent on buffer molarity from 25 to 75 mM and on pH from 5.8 to 7.4. Enzyme stability and activity remained unchanged in distilled water. The initial reaction rate increased from 12.8 to 23.8 g acrylamide/g dry cell/h, but NHase half-life dropped from 33 to roughly 7 h when temperature was varied from 4°C to 10°C. The addition of butyric acid up to 20 mM did not improve enzyme operational stability, and largely reduced (94%) enzyme activity. Acrylonitrile caused an irreversible damage to NHase activity. High acrylonitrile conversion (86%) was attained using 0.23 mg cells/ml in a continuously operating reactor.  相似文献   


5.
The kinetics and equilibria of complex formation by Ga(III) with NCS in aqueous solution have been measured over a range of acidities and temperatures, the contributing paths to the reaction resolved, and their rate constants and activation parameters determined. The hydrolysis equilibria required to carry out this resolution of kinetic behaviour have also been measured.

Unlike the other reported complexation reactions of Ga(III) in aqueous solution, the separate reaction pathways can be assigned with no ambiguity. At 25 °C and ionic strength 0.5 M, the observed forward rate constant for the complex formation is described by {k1 + k2K1h/[H+] + k3K1hK2h/[H+]2} M−1 s−1. For these conditions, the first and second successive hydrolysis constants of Ga(H2O)63+ are given by pK1h = 3.69 ± 0.01 and pK2h = 3.74 ± 0.04. The rate constants corresponding to the reactions of the species Ga(H2O)63+, Ga(H2O)5(OH)2+ and Ga(H2O)4(OH)2+ with NCS are k1 = 57 ± 4 M−1 −1, k2 = (1.08 ± 0.01) × 105 M−1 s−1 and k3 = 3 × 106 M−1 s−1 respectively. The complexation equilibrium quotient [GaNCS2+]/([Ga3+][NCS]) has been independently determined by spectrophotometric titration to be 20.8 ± 0.3 M−1 at 25 °C and ionic strength 0.5 M.

These kinetic results lead to an interpretation of the data, and a reinterpretation of other data for aquo-Ga(III) complex formation kinetics from the literature which support the assignment of a dissociative interchange mechanism for these reactions rather than the associative activation mode sometimes proposed.  相似文献   


6.
The thermal stability of a highly purified preparation of D-amino acid oxidase from Trigonopsis variabilis (TvDAO), which does not show microheterogeneity due to the partial oxidation of Cys-108, was studied based on dependence of temperature (20-60°C) and protein concentration (5-100 µmol L-1). The time courses of loss of enzyme activity in 100 mmol L-1 potassium phosphate buffer, pH 8.0, are well described by a formal kinetic mechanism in which two parallel denaturation processes, partial thermal unfolding and dissociation of the FAD cofactor, combine to yield the overall inactivation rate. Estimates from global fitting of the data revealed that the first-order rate constant of the unfolding reaction (k a) increased 104-fold in response to an increase in temperature from 20 to 60°C. The rate constants of FAD release (k b) and binding (k -b) as well as the irreversible aggregation of the apo-enzyme (k agg) were less sensitive to changes in temperature, their activation energy (E a) being about 52 kJ mol-1 in comparison with an E a value of 185 kJ mol-1 for k a. The rate-determining step of TvDAO inactivation switched from FAD dissociation to unfolding at high temperatures. The model adequately described the effect of protein concentration on inactivation kinetics. Its predictions regarding the extent of FAD release and aggregation during thermal denaturation were confirmed by experiments. TvDAO is shown to contain two highly reactive cysteines per protein subunit whose modification with 5,5'-dithio-bis (2-nitrobenzoic acid) was accompanied by inactivation. Dithiothreitol (1 mmol L-1) enhanced up to 10-fold the recovery of enzyme activity during ion exchange chromatography of technical-grade TvDAO. However, it did not stabilize TvDAO at all temperatures and protein concentrations, suggesting that deactivation of cysteines was not responsible for thermal denaturation.  相似文献   

7.
The hydrolysis of 2,4-dinitrophenylphosphate (DNPP) to orthophosphate and 2,4-dinitrophenolate (DNP) is accelerated in the presence of excess tn2Co(H2O)23+ or trpnCo(H2O)23+ at rates which maximize at pHs close to those at which the hydroxoaquatetraaminecobalt(III) complex concentrations peak (tn2, pH 6.4; trpn, pH 6.0; tn = trimethylenediamine; trpn = 3,3′,3″-triaminotripropylamine). For dilute DNPP solutions (10−4 M) the hydrolysis rates (25°C, 0.50 M NaClO4) increase with increasing Co/DNPP ratio in ways that are qualitatively as well as quantitatively different for the two systems (trpn: steady increase moving toward rate saturation, higher rates; tn2: ‘S’-shaped curve with very low rates at low ratios, lower rates compared to trpn for comparable ratios). For the trpn system the results are interpreted on the basis of pre-equilibrium formation of the 1:1 monodentate-DNPP cobalt complex by substitution of the labile water on cobalt, and rate-determining attack by the cis-coordinated hydroxide on the phosphorus center to affect hydrolysis. For the tn2 system the main path to hydrolysis is through a 2:1 cobalt to DNPP complex in which attack by a cis-coordinated hydroxide is again involved. The more complex rate behavior and the slower hydrolysis rates observed for tn2 system result from the formation of cis and trans isomers in which trans arrangements of coordinated DNPP and hydroxide leave the latter unavailable to participate in intramolecular hydrolysis. Computer fitting of the observed rate data provides values of equilibrium and rate constants for the two systems. Detailed mechanistic schemes are proposed. For the trpn system at pH 6.0 and a 25:1 cobalt to DNPP ratio (5 × 10−5 M DNPP) the observed acceleration over hydrolysis in the absence of the cobalt complex is 3 × 103; the calculated specific rate constant for hydrolysis in the reactive 1:1 complex (k 0.2 s−1) represents an acceleration over the unpromoted rate of 3 × 104.  相似文献   

8.
Electron self-exchange in solutions of the ‘blue’ copper protein plastocyanin is catalysed by the redox-inert multivalent cations Mg2+ or Co(NH3)3+6. Measurements of specific 1H-NMR line broadening with 50% reduced solutions in the presence of these cations show that electron exchange proceeds through encounters of cation-protein complexes which dissociate at high ionic strength. In the presence of 8mM (5 equivalents/total protein) Co(NH3)3+6, with 10 mM cacodylate (pH*6.0) as background electrolyte, the bimolecular rate constant at 25°C is 7 × 104 M−1·s−1. For comparison, the ‘electrostatically screened’ rate constant measured in 0.1 M KCl in the absence of added multivalent cations is ˜ 4 × 103 M1·s−1.

Plastocyanin Electron self-exchange NMR Protein-protein interaction Multivalent cation Blue copper protein  相似文献   


9.
The reaction between a cytochrome oxidase from Pseudomonas aeruginosa and oxygen has been studied by a rapid mixing technique. The data indicate that the heme d1 moiety of the ascorbate-reduced enzyme is oxidized faster than the heme c component. The oxidation of heme d1 is accurately second order with respect to oxygen and has a rate constant of 5.7 · 104 M−1 · s−1 at 20 °C. The oxidation of the heme c has a first-order rate constant of about 8 s−1 at infinite concentration of O2. The results indicate that the rate-limiting step is the internal transfer of electrons from heme c to heme d1. These more rapid reactions are followed by more complicated but smaller absorbance changes whose origin is still not clear.

The reaction of ascorbate-reduced oxidase with CO has also been studied and is second order with a rate constant of 1.8 · 104 M−1 · s−1. The initial reaction with CO is followed by a slower reaction of significantly less magnitude. The equilibrium constant for the reaction with CO, calculated as a dissociation constant from titrimetric experiments with dithionite-reduced oxidase, is about 2.3 · 10−6 M. From these data a rate constant of 0.041 s−1 can be calculated for the dissociation of CO from the enzyme.  相似文献   


10.
Thor Arnason  John Sinclair 《BBA》1976,430(3):517-523
The modulated oxygen polarograph has been used to study the rate-determining steps of photosynthetic oxygen evolution in spinach chloroplasts. The rate constant, k, of the reaction has a value of 218±10 (S.E.) s−1 at 23 °C and an activation energy of 7±2 (S.E.) kcal · mol−1. A kinetic isotope experiment indicated that this step is probably not the water-splitting reaction. These findings resemble previous results with the unicellular alga Chlorella (Sinclair, J. and Arnason, T. (1974) Biochim. Biophys. Acta 368, 393–400). In other experiments we changed the pH, O2 concentration and osmolarity of the medium, and treated the chloroplasts with 1 mM NH4Cl without detecting any significant change in k. These results suggest that the step is irreversible. However, a significantly lower value of k, 110±20 (S.E.) s−1 was obtained when all salts except 1 mM MgCl2 were removed from the medium bathing the chloroplasts.  相似文献   

11.
The reaction of peroxynitrous acid with monohydroascorbate, over the concentration range of 250 μM to 50 mM of monohydroascorbate at pH 5.8 and at 25°C, was reinvestigated and the rate constant of the reaction found to be much higher than reported earlier (Bartlett, D.; Church, D. F.; Bounds, P. L.; Koppenol, W. H. The kinetics of oxidation of L-ascorbic acid by peroxynitrite. Free Radic. Biol. Med. 18:85–92; 1995; Squadrito, G. L.; Jin, X.; Pryor, W. A. Stopped-flow kinetics of the reaction of ascorbic acid with peroxynitrite. Arch. Biochem. Biophys. 322:53–59; 1995). The new rate constants at pH 5.8 are k1 = 1 × 106 M−1 s−1 and k−1 = 500 s−1 for 25°C and k1 = 1.5 × 106 M−1 s−1 and k−1 = 1 × 103 s−1 for 37°C. These values indicate that even at low monohydroascorbate concentrations most of peroxynitrous acid forms an adduct with this antioxidant. The mechanism of the reaction involves formation of an intermediate, which decays to a second intermediate with an absorption maximum at 345 nm. At low monohydroascorbate concentrations, the second intermediate decays to nitrate and monohydroascorbate, while at monohydroascorbate concentrations greater than 4 mM, this second intermediate reacts with a second monohydroascorbate to form nitrite, dehydroascorbate, and monohydroascorbate. EPR experiments indicate that the yield of the ascorbyl radical is 0.24% relative to the initial peroxynitrous acid concentration, and that this small amount of ascorbyl radicals is formed concomitantly with the decrease of the absorption at 345 nm. Thus, the ascorbyl radical is not a primary reaction product. Under the conditions of these experiments, no homolysis of peroxynitrous acid to nitrogen dioxide and hydroxyl radical was observed. Aside from monohydroascorbate's ability to “repair” oxidatively modified biomolecules, it may play a role as scavenger of peroxynitrous acid.  相似文献   

12.
In addition to the (Na++K+)ATPase another P-ATPase, the ouabain-insensitive Na+-ATPase has been observed in several tissues. In the present paper, the effects of ligands, such as Mg2+, MgATP and furosemide on the Na+-ATPase and its modulation by pH were studied in the proximal renal tubule of pig. The principal kinetics parameters of the Na+-ATPase at pH 7.0 are: (a) K0.5 for Na+=8.9±2.2 mM; (b) K0.5 for MgATP=1.8±0.4 mM; (c) two sites for free Mg2+: one stimulatory (K0.5=0.20±0.06 mM) and other inhibitory (I0.5=1.1±0.4 mM); and (d) I0.5 for furosemide=1.1±0.2 mM. Acidification of the reaction medium to pH 6.2 decreases the apparent affinity for Na+ (K0.5=19.5±0.4) and MgATP (K0.5=3.4±0.3 mM) but increases the apparent affinity for furosemide (0.18±0.02 mM) and Mg2+ (0.05±0.02 mM). Alkalization of the reaction medium to pH 7.8 decreases the apparent affinity for Na+ (K0.5=18.7±1.5 mM) and furosemide (I0.5=3.04±0.57 mM) but does not change the apparent affinity to MgATP and Mg2+. The data presented in this paper indicate that the modulation of the Na+-ATPase by pH is the result of different modifications in several steps of its catalytical cycle. Furthermore, they suggest that changes in the concentration of natural ligands such as Mg2+ and MgATP complex may play an important role in the Na+-ATPase physiological regulatory mechanisms.  相似文献   

13.
John Sinclair  Thor Arnason 《BBA》1974,368(3):393-400
The modulated polarographic technique of O2 detection was applied to Chlorella to study the rate-limiting thermal reaction between Photosystem II and O2 evolution. From an analysis of the operation of the polarograph at different frequencies, it was concluded that a first order thermal reaction of rate constant 305±20 (S.E.) s−1 was consistent with the results of 22 °C. When the algae were successively studied in solutions made up with 2H2O and H2O, a kinetic isotopic effect for the rate constant of 1.29±0.05 (S.E.) was found. This suggests that the rate limiting step does not involve the breaking of the O-H bond in water. A temperature study of the rate constant indicated an activation energy of 5.9±0.5 (S.E.) kcal·mole−1 and an entropy of activation of −25 cal·degree−1·mole−1. The linearity of the Arrhenius plot between 8 and 42 °C demonstrated that only one reaction was rate-limiting over this temperature range.  相似文献   

14.
The ester cleavage of R- and S-isomers N-CBZ-leucine p-nitrophenyl ester intermolecularly catalyzed by R- (a) and S-stereoisomers (b) of the Pd(II) metallacycle [Pd(C6H4C*HMeNMe2)Cl(py)] (3) follows the rate expression kobs = ko + kcat [3], where the rate constants kcat equal 25.8 ± 0.4 and 7.6 ± 0.5 dm3 mol−1 s−1 for the S- and R-ester, respectively, in the case of 3a, but are 5.7 ± 0.6 and 26.7 ± 0.5 dm3 mol−1 s−1 for the S- and R-ester, respectively, in the case of 3b (pH 6.23 and 25°C). Thus, the best catalysis occurs when the asymmetric carbons of the leucine ester and Pd(II) complex are R and S, or S and R configured, respectively. Molecular modeling suggests that the stereoselection results from the spatial interaction between the CH2CHMe2 radical of the ester and the -methyl group of 3. A hydrophobic/stacking contact between the leaving 4-nitrophenolate and the coordinated pyridine also seems to play a role. Less efficient intramolecular enantioselection was observed for the hydrolysis of N-t-BOC-S-metthionine p-nitrophenyl ester with R- and S-enantiomers of [Pd(C6H4C*HMeNMe2)Cl] coordinated to sulfur.  相似文献   

15.
1. Fluoride is a mixed-type inhibitor of the cytochrome c oxidase activity with a Ki for the free enzyme of 10 mM and a Ki for the cytochrome c-complexed enzyme of 35 mM.

2. Fluoride shifts the γ-band of the enzyme from 423 to 421 nm and the -band from 597 to 598 nm. The difference spectrum (oxidized enzyme in the presence of fluoride minus oxidized enzyme) has peaks at 400, 453, 482, 605 and 638 nm and troughs at 430, 520, 552 and 674 nm. The changes in absorbance are small (about 3% at absorbance maxima) with respect to those of other hemoproteins.

3. On addition of fluoride to isolated cytochrome c oxidase 3 reactions can be distinguished: (I) a bimolecular binding reaction (Kon = 4 M−1 · s−1 and koff = 2.9 · 10−2s−1 at 25 °C, pH 7.4) contributing at 638 nm and 430 nm; (II) a first-order reaction (k = 2.4 · 10−2) s−1 at 22 °C, pH 7.2) visible mainly at 430 nm and (III) a very slow reaction with a half-time in the order of 10 min.

4. The spectroscopic dissociation constants for the fluoride binding, determined from Hill plots using the absorbance changes at 638 and 430 nm, are similar (7 and 10 mM, respectively, at 22 °C, pH 7.2).

5. A mechanism for the reaction is discussed in which the bimolecular binding reaction is followed by a conformational change of the enzyme-fluoride complex.  相似文献   


16.
A kinetic model has been used to estimate the rate constant for the reaction of superoxide (O2/OOH) with the superoxide spin adduct of 5.5-dimethylpyrroline-N-oxide. DMPO/OOH. This rate constant is estimated to be 4.9 (± 2.2) × 106 M-1 s-1, pH 7.4 and 25°C.  相似文献   

17.
Alpha-Chymotrypsin was found to show a 119% increase in activity after three phase partitioning. The kcat/Km of the partitioned enzyme (TPP-C) for hydrolysis of Bz-Tyr-OEt in aqueous medium at 25°C was found to be 48.3×104 mM-1 min-1 as compared to the corresponding value of 17.7×104 mM-1 min-1 for the untreated control (C). The λmax of the fluorescence emission spectrum of TPP-C showed 178% increase in the quantum yield when compared to C. TPP-C showed a 2.94 and 3.58 fold increase (as compared to C) in initial rates for formation of the ester Ac-Phe-OEt (from Ac-Phe and ethanol) in low water containing toluene and n-octane, respectively. It was found that TPP-C also showed the phenomenon of pH memory. At 5% (v v-1) water (in t-amyl alcohol), while no esterification was observed with C, TPP-C still showed significant level of esterification activity.  相似文献   

18.
Somatostatin binding to guinea pig pancreatic acinar cell plasma membranes was characterized with an iodinated stable analog of somatostatin 28 (S28): 125I-[Leu8, DTrp22,Tyr25] S28. The binding was highly dependent on calcium ions. In 0.2 mM free Ca2+ medium, binding at 37°C was saturable, slowly reversible and exhibited a single class of high affinity binding sites (KD=0.05±0.01 nM, Bmax=157±33 fmol/mg protein). Dissociation of bound radioactivity occurred with biphasic kinetics. Rate of dissociation increased when dissociation was measured at a time before equilibrium binding was reached. In 30 nM free Ca2+ medium, binding affinity and maximal binding capacity were decreased by about 4-fold. Decreasing calcium concentrations increased the amount of rapidly dissociating form of the receptor. Somatostatin 14 antagonist, Des AA1,2[AzaAla4–5,DTrp8,Phe12–13]-somatostatin was active at the membrane level in inhibiting the binding. We conclude that using 125I-[Leu8,DTrp22,Tyr25]S28 as radioligand allows us to characterize a population of specific somatostatin receptors which are not different from those we previously described with the radioligand 125I-[Tyr11]-somatostatin. Somatostatin receptors could exist in two interconvertible forms. Calcium ions are an essential component in the regulation of the conformational change of somatostatin receptors.  相似文献   

19.
Trypanosoma cruzi trypanothione reductase (TR) was irreversibly inhibited by peroxidase/H2O2/phenothiazine (PTZ) systems. TR inactivation depended on (a) time of incubation with the phenothiazine system; (b) the peroxidase nature and (c) the PTZ structure and concentration. With the most effective systems, TR inactivation kinetics were biphasic, with a relatively fast initial phase during which about 75% of the enzyme activity was lost, followed by a slower phase leading to total enzyme inactivation. GSH prevented TR inactivation by the peroxidase/H2O2/PTZ systems. Production of PTZ cation radicals by PTZ peroxidation was essential for TR inactivation. Horseradish peroxidase, leukocyte myeloperoxidase (MPO) and the pseudo-peroxidase myoglobin (Mb) were effective catalysts of PTZ production. Promazine, thioridazine, chlorpromazine, propionylpromazine prochlorperazine, perphenazine and trimeprazine were effective constituents of the HRP/H2O2/PTZ system. The presence of substituents at the PTZ nucleus position 2 exerted significant influence on PTZ activity, as shown by the different effects of 2-trifluoromethyl and 2-H or 2-chlorophenothiazines. The PTZ cation radicals disproportionation regenerated the non-radical PTZ molecule and produced the PTZ sulfoxide that was inactive on TR. Thiol compounds including GSH interacted with PTZ cation radicals transferring an electron from the sulfide anion to the PTZ, thus nullifying the PTZ biological and chemical activities.  相似文献   

20.
H.F. Kauffman  B.F. Van Gelder 《BBA》1973,314(3):276-283
1. Cyanide causes a slow disappearance of the oxidized band (648 nm) of cytochrome d in particles of Azotobacter vinelandii and inhibits the appearance of the reduced band (631 nm). No effect of cyanide is found on the reduced band of cytochrome d.

2. The kinetics of the disappearance of the 648-nm band of cytochrome d with excess cyanide deviates from first-order kinetics at lower temperatures (22 °C) indicating that at least two conformations of the enzyme are involved. At higher temperatures (32 °C) the observed kinetics of the cyanide reaction are first order with a kon = 0.7 M−1·s−1 and with an estimated koff of approximately 5·10−5 s−1.

3. The value of the koff (7·10−4−14·10−4 s−1 at 32 °C) determined from the rate of reduction of cyanocytochrome d by Na2S2O4 or NADH is one order of magnitude larger than the koff value found when the enzyme is in its oxidized state.

4. No effect of cyanide is found on the spectrum of cytochrome a1.  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