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 390 毫秒
1.
The catalytic activity of papain in the synthesis of Z-Gly-Phe-NH2 in tert-butanol has been studied in the presence of solid-state acid–base buffers (acids and their sodium salts). All buffer pairs tested reduced the reaction rate compared with the control, particularly the most acidic and basic (assessed by either aqueous pKa or the response of an organic phase indicator). The highest rates, close to the control, were found with glutamic acid/glutamate-Na, PIPES/PIPES-Na and NaH2PO4/Na2HPO4. However, these pairs were unable to erase the pH memory phenomenon, or to overcome the effect of spiking with acetic acid. Hence, at least these buffers do not seem to be able to affect the protonation state and catalytic activity of papain. In the last aqueous solution before drying, the presence of activating agents (cysteine plus EDTA) was more important than buffer ions.  相似文献   

2.
We report on the performance of cross-linked enzyme microcrystals (CLECs) of subtilisin Carlsberg in supercritical fluids (SC-fluids). The catalytic activity of CLECs in SC-ethane was found to be 2- to 10-fold greater than in hexane under the same conditions, using CLECs dried by propanol washing. Air-dried CLECs and lyophilized powders showed much lower activities, reflecting the same hydration hysteresis effects as in organic solvents. Reaction rates were much lower in SC-CO(2), especially at higher water activity, probably as a result of acid-base effects of carbonic acid on the enzyme.  相似文献   

3.
Enzyme performance is often impaired in supercritical carbon dioxide. We were able to enhance enzyme activity in this medium via changes in acid-base conditions by using ion-exchange materials (solid H(+)/Na(+) buffer pairs and a zeolite), which were selected on the basis of the response of an organosoluble acid-base indicator. The concentration of ion-exchange materials had an important effect on the catalytic activity of subtilisin Carlsberg cross-linked enzyme crystals (CLECs), and this was related to the protonation and hydration states of the enzyme. The buffer Na(2)CO(3)/NaHCO(3) gave the highest enhancement in enzyme activity (by a factor of 54), probably as a result of its high basicity and capacity to counteract the deleterious effect of carbonic acid to a greater extent than the other materials tested.  相似文献   

4.
Recently, we reported a new high-activity biocatalyst for use in organic media termed protein-coated microcrystals (PCMC) (Kreiner et al. [2001] Chem Commun 12:1096-1097). These novel particles consist of water-soluble micron-sized crystalline particles coated with the given biocatalyst(s) and are prepared in a one-step rapid dehydration process. In this study we extended the choice of immobilisation matrix from a simple inorganic salt, K(2)SO(4), to other compounds, both inorganic and zwitterionic, that act as solid-state buffers for biocatalysis in organic media. The catalytic activity of serine proteases subtilisin Carlsberg (SC) and alpha-chymotrypsin (CT) were significantly increased when coated onto the surface of solid-state buffers, as measured in acetonitrile/1wt% H(2)O. SC-PCMC with both organic and inorganic buffer carriers (Na-AMPSO, Na(2)CO(3), and NaHCO(3)) showed a 3-fold greater activity than that observed when using the unbuffered system (PCMC-SC/K(2)SO(4)). In comparison with freeze-dried preparations, this represents an approximately 3,000-fold increase in catalytic activity. Importantly, there is no improvement in catalytic activity upon external addition of any of the solid-state buffers to the reaction mixture. When acting in a solid-state buffer capacity, good buffering capacity was observed with SC-PCMC (3 wt% protein loading) prepared from a 1:1 mixture of AMPSO and AMPSO-Na. Alternatively, increasing the amount of solid-state buffer in the system allows improvement of the buffering. This can be achieved either by decreasing the protein loading of the SC/Na-AMPSO-PCMC or by addition of further external solid-state buffer to the reaction mixture. The catalytic activity of lipase-PCMC prepared from solid-state buffers was found less responsive to immobilisation.  相似文献   

5.
Cross-linked enzyme crystals(CLECs) of subtilisin display the improved thermostability in organic solvents, compared to free subtilisin. CLECs are more stable than the free enzyme in octane with a half life of 200 days at 45°C, while that of free enzyme is 5.4 days. CLECs in octane is more stable than in acetonitrile.  相似文献   

6.
The activity and the enantioselectivity of cross-linked enzyme crystals (CLECs) of subtilisin in the transesterification between N-acetyl-l-phenylalanine ethyl ester and n-propanol have been examined in various organic solvents. The activity of CLECs of subtilisin in decane was 780 times greater than that in triethylamine. CLECs of subtilisin preferred l-enantiomer in the transesterification between N-acetyl-phenylalanine ethyl ester and n-propanol, and the (kcat/KM)L/(kcat/KM)D ratio was 20 000 in cyclohexane.  相似文献   

7.
The proteolytic extract obtained from the latex of Funastrum clausum (Jacq.) Schlechter (Asclepiadaceae), a South American climbing plant, was assayed as a novel catalyst for peptide synthesis and compared with commercial papain under the same conditions. After immobilization on polyamide, the synthesis of the bitter peptide precursor Z-Ala-Phe-OMe was performed and different conditions were tried. Acetonitrile and ethyl acetate with low water content were tested as organic solvents. Equilibrium- and kinetically-controlled synthesis were tried by using either Z-Ala-OH or Z-Ala-OMe as acyl donors, respectively. The best conditions for the synthesis of the desired product varied according to the catalyst used. For papain, thermodynamic control in acetonitrile (aw  0.12) in the presence of triethylamine (TEA) or boric acid–borate buffer (40 mM), and equilibrium- and kinetic-controlled synthesis in ethyl acetate (aw  0.75) proved to be the best conditions. The thermodynamic control in either acetonitrile with aw  0.12 (40 mM TEA or Na2CO3) or ethyl acetate (aw  0.75) were the best conditions found for funastrain. In all cases, the formation of oligopeptides up to three Phe was observed. The proteolytic extract of F. clausum latex showed more selectivity than papain towards the conversion to Z-Ala-Phe-OMe leading to less proportion of oligopeptides.  相似文献   

8.
The first studies on the operational stability of cross-linked enzyme crystals (CLECs) in organic media are described. Although these catalysts display high initial specific activity, they inactivate rapidly, losing more than 50% of the initial activity within the first 4 h under continuous flow. Furthermore, the inactivation is not reversible when returned to an aqueous medium. The same rapid inactivation occurs with adsorbed protease preparations that show similar high initial specific activity (propanol-rinsed enzyme preparations (PREPs) of subtilisin and alpha-chymotrypsin).  相似文献   

9.
Amino groups in alpha-chymotrypsin were reacted with pyromellitic anhydride, introducing 17 to 32 additional carboxyl groups. This modification causes a major change in the water adsorption isotherm of the lyophilized protein powder. Little water is bound by the modified enzyme at water activity (aw) below 0.35, but it shows increased water binding at aw over 0.5. This correlates with a similar change in the aw dependence of the catalytic activity of the enzyme powder suspended in hexane, with a much steeper increase in activity of the modified chymotrypsin.  相似文献   

10.
Gay Goodman  John S. Leigh  Jr. 《BBA》1987,890(3):360-367
The electron-spin relaxation rates of the two species of cytochrome a3+3-azide found in the azide compound of bovine-heart cytochrome oxidase were measured by progressive microwave saturation at T = 10 K. It has been shown previously that Cyt a+33-azide gives rise to two distinct EPR resonances, depending upon the oxidation state of Cyt a. When Cyt a is ferrous, Cyt a3+3-azide has g = 2.88, 2.19 and 1.64; upon oxidation of Cyt a, the a3+3-azide g-values become g = 2.77, 2.18, and 1.74 (Goodman, G. (1984) J. Biol. Chem. 259, 15094–15099). The relaxation effect of Cyt a on Cyt a3 could be measured as the difference in microwave field saturation parameter H1/2 between the g = 2.77 and g = 2.88 species. For each signal the spin-lattice relaxation time T1 was determined from H1/2 using the transverse relaxation time T2. The value of T2 at 10 K was extrapolated from a plot of line-width vs. temperature at higher temperature. The dipolar contribution to T1 was related to the Cyt a-Cyt a3 spin-spin distance utilizing available information on the relative orientation of Cyt a3-azide and Cyt a (Erecinska, M., Wilson, D.F. and Blasie, J.K. (1979) Biochim. Biophys. Acta 545, 352–364). By taking into account the relaxation parameters for both gx and gz components of the Cyt a3-azide g-tensor, the angle between the gz components of the Cyt a and Cyt a3g-tensors was determined to be between 0 and 18°, and the Cyt a-Cyt a3 spin-spin distance was found to be 19 ± 8 Å.  相似文献   

11.
The thermophilic blue-green alga Synechococcus lividus was grown at 38 and 55°C. The reaction center chlorophyll-protein complexes (CP) of Photosystem (PS I) and PS II, CP aI and CP aII, were isolated by sodium dodecyl sulphate (SDS) polyacrylamide gel electrophoresis at 4°C. SDS solubilization of thylakoids was performed in the temperature range 0–65°C. The low-temperature absorption and fluorescence emission spectral properties of the isolated chlorophyll-protein complexes were analyzed. Only traces of CP aI were solubilized at temperatures below the lipid phase transition temperature. Instead, a minor PS I component, CP aI, was obtained that had absorption and fluorescence characteristics similar to those of CP aI. CP aI had a slightly lower mobility than CP aI in SDS-polyacrylamide gel electrophoresis. The amount of CP aI in the gel scan profile increased dramatically when solubilization was carried out above the phase transition temperatures, but started to decrease above 60°C. CP aII, on the other hand, could be efficiently extracted even at 0°C and was stable in the scan profile up to extraction temperatures of 30–40°C. Low-temperature absorption and fluorescence emission spectra were typical for CP aI and CP aII and no specific effects of the two growth temperatures on these properties were observed. The phase transition temperature was considered to be critical for the solubilization of CP aI, either because of the difficulties of SDS (especially as it forms micelles at low temperatures) in penetrating the solidified membrane lipids at temperatures below that of the phase transition or because the CP aI monomers of the PS I antennae are so strongly bound to each other that they cannot be dissociated by SDS before thermal agitation has reached a certain level that is achieved above the phase transition temperature. We consider both the difficulties in solubilizing CP aI at sub-transition temperatures and the heat stability of the two complexes as adaptations which enable Synechococcus to grow under extreme high-temperature regimes.  相似文献   

12.
The molecular structure and physicochemical properties of acid–alcohol treated maize and potato starches (0.36% HCl in methanol at 25 °C for 1–15 days) were investigated. The yields of the modified starches were ranging from 91 to 100%. The average granule size of modified starches decreased slightly. The solubility of starches increased with the increase of treatment time, and the pasting properties confirmed the high solubility of modified starches. The gelatinization temperatures and range of gelatinization increased with the increase of treatment time except To (onset temperature) of maize starch. Molecular structures of modified starches suggested the degradation of starches occurred mostly within the first 5 days of treatment, and degradation rate of potato starch was higher than maize starch both in amylopectin and in amylose. Maize starch was found less susceptible to acid–alcohol degradation than potato starch.  相似文献   

13.
The water activity dependence of lipase kinetics in organic media was evaluated using lipases from Rhizopus oryzae and Candida rugosa immobilised on polypropene EP-100. The conversion studied was the transesterification of ethyl decanoate to hexyl decanoate with hydrolysis to decanoic acid as competing reaction. The reactions were carried out at controlled water activity in diisopropyl ether. Substrate inhibition was observed at hexanol concentrations of 100 mM or higher. The Rhizopus lipase expressed the highest activity and the best selectivity for transesterification at the lowest water activity (aw=0.06). The Candida lipase expressed the highest transesterification/hydrolysis ratio at aw=0.11 and the highest total activity at aw=0.53. Several glycosidases previously tested under conditions similar to those used here expressed both maximal total activity and the best selectivity at water activities close to 1.0. The water activity dependence of the lipases is thus fundamentally different from that of glycosidases and it is a major part of the reason why lipases are more suited for transferase-type reactions than the glycosidases.  相似文献   

14.
Careful base and solvent optimization for catalytic amination is described. A Pd-catalyzed amination between some arylbromide and unprotected piperazine (1 equiv) was efficiently carried out with Pd/BINAP catalyst in a toluene–DBU solvent system, which is useful for the one-pot preparation of unsymmetrical piperazine through amination and in-situ N-protection. Reaction with N-BOC-piperazine was also successful in toluene–DBU or more polar NMP with Cs2CO3 as a key base. No reports have previously reported such solvent and base optimization in arylpiperazine synthesis.  相似文献   

15.
Recently, a new technique using cross-linking enzyme crystals (CLECs) was introduced in the field of enzyme technology. CLECs are solid crystalline particles which are insoluble in both water and organic solvents. Chemical cross-linking of the enzyme crystals preserves the catalytic activity even in harsh conditions such as at high temperature, at extreme pH, in organic solvents, and in the presence of proteases and radical chemicals. CLECs of cyclodextrin glucanotransferase (CGTase) could be useful biocatalysts because they were stable at elevated temperature, in organic solvents, and in the presence of enzyme inactivation surfactant. They also maintained their activity against protein-digesting enzyme.  相似文献   

16.
Growth of Gibberella fujikuroi on Amberlite, an inert support, and gibberellic acid (GA3) production was studied in glass columns under different conditions of temperature and water activity (aw). Maximum biomass concentration and GA3 production were respectively 40 (mg/g inert support) and 0.73 (mg/g inert support). While high specific growth rates were obtained, low initial nitrogen resulted in low biomass concentrations. Maximum GA3 (31°C, aw=0.985) was not produced by the maximum concentration of biomass (25°C, aw=0.992). Peaks in the rate curves of either outlet gas, CO2 or O2, occurred on exhaustion of urea indicating, for future works, just when to feed the culture additional nitrogen.  相似文献   

17.
The reactivity of horseradish peroxidase (HRP) with water insoluble phenolic compounds has been studied in 1-butyl-3-methylimidazolium tetrafluoroborate ([BMIM][BF4])/water mixtures. The enzyme retained some catalytic activity up to 90% ionic liquid in water at 25 °C only at pH values higher than 9.0. Activity steadily decreased towards neutral and acidic conditions, as judged by 4-aminoantypirin/phenol activity tests. Inhibition of horseradish peroxidase under neutral acidic condition was due to the binding of fluoride anions released from tetrafluoroborate anion to the heme iron as demonstrated by the sharp UV–visible absorption transition diagnostic of the conversion from a five coordinated to a six coordinated high spin ferric heme iron. Thus, reactions with water insoluble phenols were carried out under alkaline reaction conditions and 75% [BMIM][BF4]/water mixture. Under these conditions, the distribution of the reaction products was much narrower with respect to that observed in aqueous buffers or water/dimethylformamide or water/dimethylsulfoxide mixtures, and polymeric species other than dimers were not observed. Technical scale preparations of a novel 4-phenylphenol ortho dimer [2,2′-bi-(4-phenylphenol)] with a high yield of the desired product were obtained.  相似文献   

18.
Salt hydrates very frequently are utilized as in situ water activity buffers in reaction mixtures of enzymes in nonaqueous media. In addition to buffering water activity, there is evidence that salt hydrates also often affect initial rates in other ways. This has been generally overlooked or thought to be related to water transfer effects. Here we show that salt hydrates can have important acid-base effects on enzymes in nonaqueous media. We performed transesterification reactions in n-hexane and in supercritical ethane catalyzed by cross-linked crystals of subtilisin, differing in the method used to set a(W), and confirmed that the presence of salt hydrate pairs significantly affected the catalytic performance of the enzyme. However, in the presence of a solid-state acid-base buffer, salt hydrates had no effect on enzymatic activity. Direct evidence for the acid-base effects of salt hydrates was obtained by testing their effect on the protonation state of an organo-soluble H(+)/Na(+) indicator. The four salt hydrate pairs tested affected the indicator to very different extents. By promoting the exchange of H(+) for Na(+), salt hydrates will tend to affect the ionization state of acidic residues in the protein and, hence, enzymatic activity. In fact, salt hydrates were able to affect the pH memory of subtilisin lyophilized from different aqueous pHs, bringing about up to 20-fold enhancements and up to 5-fold decreases in catalytic activity. The possibility of such acid-base effects need to be considered in all experiments using salt hydrates to control water activity.  相似文献   

19.
(1) The reaction of the resting form of oxidised cytochrome c oxidase from ox heart with dithionite has been studied in the presence and absence of cyanide. In both cases, cytochrome a reduction in 0.1 M phosphate (pH 7) occurs at a rate of 8.2 · 104 M−1 · s−1. In the absence of cyanide, ferrocytochrome a3 appears at a rate (kobs) of 0.016 s−1. Ferricytochrome a3 maintains its 418 nm Soret maximum until reduced. The rate of a3 reduction is independent of dithionite concentration over a range 0.9 mM–131 mM. In the presence or cyanide, visible and EPR spectral changes indicate the formation of a ferric a3/cyanide complex occurs at the same rate as a3 reduction in the absence of cyanide. A g = 3.6 signal appears at the same time as the decay of a g = 6 signal. No EPR signals which could be attributed to copper in any significant amounts could be detected after dithionite addition, either in the presence or absence of cyanide. (2) Addition of dithionite to cytochrome oxidase at various times following induction of turnover with ascorbate/TMPD, results in a biphasic reduction of cytochrome a3 with an increasing proportion of the fast phase of reduction occurring after longer turnover times. At the same time, the predominant steady state species of ferri-cytochrome a3 shifts from high to low spin and the steady-state level of reduction of cytochrome a drops indicating a shift in population of the enzyme molecules to a species with fast turnover. In the final activated form, oxygen is not required for fast internal electron transfer to cytochrome a3. In addition, oxygen does not induce further electron uptake in samples of resting cytochrome oxidase reduced under anaerobic conditions in the presence of cyanide. Both findings are contrary to predictions of certain O-loop types of mechanism for proton translocation. (3) A measurement of electron entry into the resting form of cytochrome oxidase in the presence of cyanide, using TMPD or cytochrome c under anaerobic conditions, shows that three electrons per oxidase enter below a redox potential of around +200 mV. An initial fast entry of two electrons is followed by a slow (kobs ≈ 0.02 s) entry of a third electron. Above +200 mV, the number of electrons taken up in the initial fast phase drops as a redox center (presumably CuA) titrates with an apparent mid-point potential of +240 mV. The slow phase of reduction remains at the more positive redox values. (4) The results are interpreted in terms of an initial fast reduction of cytochrome a (and CuA at redox values more negative than +240 mV) followed by a slow reduction of CuB. CuB reduction is proposed to spin-uncouple cytochrome a3 to form a cyanide sensitive center, and trigger a conformational change to an activated form of the enzyme with faster intramolecular electron transfer.  相似文献   

20.
The individual and interactive effects of temperature, pH, NaCl, and aw on the proteolytic and lipolytic activities of Lactobacillus delbrueckii subsp. bulgaricus B397, Lactococcus lactis subsp. lactis T12, and Lb. plantarum 2739 were studied by quadratic response surface methodology. The effects on enzyme activities depended on the interactions among the independent variables, type of activity, substrate and, especially, species. The proteinase activity of strains B397 and T12 was affected differently by pH as individual or interactive terms depending on the type of substrate sl- or β-casein. The increase of NaCl concentration (2.5–7.5%) under cheese-like conditions had a negative effect on the proteinase activity of strain T12. The effect of NaCl was related to the corresponding decrease in aw. Aminopeptidases N and A, iminopeptidase and endopeptidase of Lc. lactis subsp. lactis T12 were strongly inhibited by pH 5–6 and NaCl concentration higher than 3.75%. The negative effects of these independent variables was in several cases enhanced by their interaction and/or by the interaction with the lowest temperatures. In contrast, the same peptidases of Lb. plantarum 2739 retained a high activity under the very hostile environmental conditions. Iminopeptidase and especially endopeptidase activities of strain 2739 were stimulated slightly by NaCl at concentrations up to 5%. Lipase/esterase activity of Lb. delbrueckii subsp. bulgaricus B397 was markedly inhibited under cheese-like conditions.  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