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 78 毫秒
1.
The initial synthetic rates of peptide Cbz-Arg-Leu-NH(2) from Cbz-Arg and Leu-NH(2) using PST-01 protease in the presence and absence of organic solvents were investigated under various conditions. The synthetic rates of Cbz-Arg-Leu-NH(2) in the presence of 50% (v/v) methanol, 50% (v/v) N,N-dimethylformamide (DMF) and 60% (v/v) dimethyl sulfoxide (DMSO) were 1.6-, 2.4-, and 5.1-times higher than that in the absence of organic solvent, respectively. The PST-01 protease was not only stable in the presence of organic solvents but also exhibited high reaction rates in the presence of methanol, DMF, and DMSO. When the Cbz-Arg concentration was lower than 60mM or the Leu-NH(2) concentration was lower than 400mM, the initial rates increased lineally with increase in their concentrations. However, the rates did not increase when the Leu-NH(2) concentration was more than 500mM. The optimum temperature and pH of the reaction were 40 degrees C and 7.0, respectively.  相似文献   

2.
The PST-01 protease is a metalloprotease that has zinc ion at the active center and is very stable in the presence of water-soluble organic solvents. The reaction rates and the equilibrium yields of the aspartame precursor N-carbobenzoxy-L-aspartyl-L-phenylalanine methyl ester (Cbz-Asp-Phe-OMe) synthesis from N-carbobenzoxy-L-aspartic acid (Cbz-Asp) and L-phenylalanine methyl ester (Phe-OMe) in the presence of water-soluble organic solvents were investigated under various conditions. Higher reaction rate and yield of Cbz-Asp-Phe-OMe were attained by the PST-01 protease when 30 mM Cbz-Asp and 500 mM Phe-OMe were used. The maximum reaction rate was obtained pH 8.0 and 37 degrees C. In the presence of dimethylsulfoxide (DMSO), glycerol, methanol, and ethylene glycol, higher reaction rates were obtained. The equilibrium yield was the highest in the presence of DMSO. The equilibrium yield of Cbz-Asp-Phe-OMe using the PST-01 protease attained 83% in the presence of 50% (v/v) DMSO.  相似文献   

3.
The membrane-bound detergent-activated ATPase from Halobacterium saccharovorum was purified at a physiological salt concentration (4 M NaCl) in the presence of nonionic detergents. The preparation contains putative subunits of 110, 71, 31, 22, and 14 kDa. The enzyme activity required high salt concentration but was not dependent on any one specific monovalent cation or any anion. The hydrolysis of ATP was nonlinear with time; the data are consistent with a kinetic model where the enzyme is irreversibly converted from an initial into a final stable form during the first few minutes of the reaction. The model thus contains a rate constant (k) for the transition and hydrolytic rates, v1 and v2, for the two forms of the enzyme. We found that this hysteretic behavior was influenced differently by various conditions and inhibitors. The constant k was smaller with Mn2+ than with Mg2+ as the divalent cation, showed negative temperature dependence, and a distinct pH optimum between 7.5 and 8.5. Thiols decreased k, but nitrate, a specific inhibitor of archaebacterial ATPases, increased it. ADP showed competitive inhibition against both the initial and the final form of the enzyme. Nitrate reversibly inhibited only the latter and in a manner dependent on whether Mn2+ or Mg2+ was used. The kinetic data suggest that all agents tested, with effects on the hydrolytic activity, seem to act at or near the catalytic site of the enzyme.  相似文献   

4.
N-(Benzyloxycarbonyl)-L-aspartyl-L-phenylalanine methyl ester (Z-AspPheOMe), a precursor of the aspartame, and N-(benzyloxycarbonyl)-L-phenylalanyl-Lphenylalanine methyl ester (Z-PhePheOMe) were synthesized from the respective amino acid derivatives with an immobilized thermolysin (EC 3.4.24.4) in ethyl acetate. Various factors affecting the synthesis of these dipeptide precursors were clarified. The initial synthetic rate was the highest at the water content of 3.5% for both reactions. The substrate concentration dependencies of the initial synthetic rate of Z-AspkPheOMe and Z-PhePheOMe with the immobilized enzyme in ethyl acetate were different from those in an aqueous buffer solution saturated with ethyl acetate but similar to those in the aqueous/organic biphasic system using the free enzyme. Particularly, the initial synthetic rate of Z-AspPhOMe increased in order higher than first order with respect to the concentration of L-phenylalanine methyl ester (PheOMe), whereas it decreased sharply with the concentration of N-(benzyloxycarbonyl)-L-aspartic acid (Z-Asp). Such kinetic behavior could be explained by regarding the inside of the immobilized enzyme as being a biphasic mode composed from the organic phase and aqueous phase where the enzymatic reaction takes place. The reaction in the aqueous/organic biphasic system using the free enzyme could be simulated by taking into consideration the partition of the substrate and the initial rate of synthesis in the aqueous buffer saturated with ethyl acetate. Based on this analysis, the rate of reaction with the immobilized enzyme in ethyl acetate could also be predicted. Z-AsPheOMe and Z-PhePheOMe were synthesized by the fed-batch method where the acid component of the substrate was intermittently added during the course of reaction and by the batch method. In the synthesis of Z-AspPheOMe, the synthetic rate and maximum yield of reaction as well as the stability of the immobilized enzyme were higher in the fed-batch reaction than those in the batch reaction. In the synthesis of Z-PhePheOMe, the results obtained by both methods were similar. (c) 1994 John Wiley & Sons, Inc.  相似文献   

5.
Enzymatic acyl-transfer reaction in organic medium competes with the hydrolytic side reaction depending on the water content. The effect of water content on aminolysis activity of -chymotrypsin for the synthesis of Bz-Tyr-Val-NH2 in acetonitrile was examined under the conditions which were devoid of the hydrolytic deacylation. Excess H-Val-NH2 (880 mM) was employed to keep the hydrolysis negligible. The aminolysis rate increased abruptly between 4 and 5% (v/v) water but a further increase in the water content did not affect the reaction rate. This suggests that water added more than 5% (v/v) does not enhance intrinsic enzyme activity but acts only as a nucleophile for the hydrolytic deacylation.  相似文献   

6.
The kinetically controlled condensation reaction of Z-Gly-Trp-Met-OR(1) (R(1): Et, Al, Cam) and H-Asp-(OR(2))-Phe-NH(2) (R(2): H, Bu(t)) catalyzed by alpha-chymotrypsin deposited onto polyamide in organic media was studied. The effect of the drying process of the enzyme-support preparation, substrate concentrations, reaction medium, acyl donor, and nucleophile structure on both enzymatic activity and pentapeptide yield was investigated. The immobilized preparation directly equilibrated at a(w) = 0.113, gave higher enzymatic activities than dried with vacuum first, and then equilibrated at a(w) = 0.113. The addition of triethylamine to the reaction medium increased dramatically the enzymatic activity. However, the pentapeptide yield was affected neither by the drying procedure nor by the addition of triethylamine. The donor ester Z-Gly-Trp-Met-OAl gave initial reaction rates 2.6 times higher than the conventional ethyl ester derivative but rendered similar yields. The best results were obtained using Z-Gly-Trp-Met-OCam as acyl-donor ester; 80% yield and initial reaction rates 4 times higher than the ethyl ester derivative. In all cases, acetonitrile containing Tris-HCl 50 mM pH 9 buffer (0.5% v/v) and triethylamine (0.5% v/v) was found to be the best reaction system. Under these conditions, it was possible to use the nucleophile H-Asp-Phe-NH(2) with beta-unprotected aspartic acid residue. In this case, 50% yield was obtained, but economic considerations could lead to select it as nucleophile. Finally, the fragment condensation reaction was carried out at gram scale, obtaining a 39% yield which included the reaction, removal of protecting groups and purification steps. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 56: 456-463, 1997.  相似文献   

7.
The enzymatic hydrolysis of a mixture of lutein diesters from Marigold flower (Tagetes erecta) was performed both in organic solvents and supercritical CO(2) (SC-CO(2)) using two commercial lipases: lipase B from Candida antarctica (Novozym 435) and the lipase from Mucor miehei (Lipozyme RM IM). Both lipases showed an unexpected dependence of initial reaction rate with the initial water activity (a(wi)) in hexane, with the highest rates of hydrolysis taking place at the lowest a(wi) of the biocatalyst particles. The same result was observed using isooctane, toluene, or SC-CO(2). It is proposed that an increase in a(wi) generates a hydrophilic microenvironment that prevents efficient partitioning of the highly hydrophobic lutein diesters to the enzyme. The critical role of water in this system has not been reported for other hydrolytic reactions in low water media. Calculations of water available for hydrolysis from isotherm analysis, Karl-Fischer titration, and substrate conversion at a(wi) = 0.13, indicate that the extent of reaction is not limited by the amount of available water. Accordingly, the enzyme that holds the largest amount of water after prehydration at the same a(wi) (0.13) will yield the greatest substrate conversion and concentration of the free lutein product. The highest conversion occurred in SC-CO(2), which opens up new opportunities to develop a combined extraction-reaction process for the environmentally benign synthesis of lutein, an important nutraceutical compound.  相似文献   

8.
1. The hydrolysis of o-nitrophenyl beta-D-glucopyranoside by the high-molecular-weight beta-glucosidase (beta-D-glucoside glucohydrolase, EC 3.2.1.21) from Botryodiplodia theobromae Pat. has been studied in the presence of added dioxan. 2. At donor saturation, the maximum rate of hydrolysis in the presence of up to 50%(v/v) dioxan was pH4.3-4.5 (pH of the buffer system in water) in McIlvaine's buffer. 3. Increasing dioxan concentrations progressively decreased the maximum rate of hydrolysis. 4. The rate of enzyme-catalysed reaction was enhanced at high donor concentrations, but inhibited at low donor concentrations in the presence of glycerol, methanol, fructose of sucrose. 5. The hydrolytic reaction was found to proceed with retention of configuration at the anomeric carbon atom. 6. The kinetics of the enzyme-catalysed process in the presence of added acceptors indicated that water was necessary for the maintenance of the active enzyme conformation apart from its acceptor function.  相似文献   

9.
Binding of streptokinase (SK) to plasminogen (Pg) induces conformational activation of the zymogen and initiates its proteolytic conversion to plasmin (Pm). The mechanism of coupling between conformational activation and Pm formation was investigated in kinetic studies. Parabolic time courses of Pg activation by SK monitored by chromogenic substrate hydrolysis had initial rates (v(1)) representing conformational activation and subsequent rates of activity increase (v(2)) corresponding to the rate of Pm generation determined by a specific discontinuous assay. The v(2) dependence on SK concentration for [Lys]Pg showed a maximum rate at a Pg to SK ratio of approximately 2:1, with inhibition at high SK concentrations. [Glu]Pg and [Lys]Pg activation showed similar kinetic behavior but much slower activation of [Glu]Pg, due to an approximately 12-fold lower affinity for SK and an approximately 20-fold lower k(cat)/K(m). Blocking lysine-binding sites on Pg inhibited SK.Pg* cleavage of [Lys]Pg to a rate comparable with that of [Glu]Pg, whereas [Glu]Pg activation was not significantly affected. The results support a kinetic mechanism in which SK activates Pg conformationally by rapid equilibrium formation of the SK.Pg* complex, followed by intermolecular cleavage of Pg to Pm by SK.Pg* and subsequent cleavage of Pg by SK.Pm. A unified model of SK-induced Pg activation suggests that generation of initial Pm by SK.Pg* acts as a self-limiting triggering mechanism to initiate production of one SK equivalent of SK.Pm, which then converts the remaining free Pg to Pm.  相似文献   

10.
Several kinds of modified chymotrypsin were prepared with water-soluble acylating reagents, and their characteristics after hydrolyzing with unmodified chymotrypsin in aqueous-N,N'-dimethylformamide (DMF) media were compared. It was found that chymotrypsin (Csin), of which a 20% amino group was modified with a benzyloxycarbonyl group (Z(20)Csin), had more favorable characteristics than unmodified chymotrypsin with regard to hydrolytic activity in an aqueous DMF media. We also investigated the Z(20)Csin-catalyzed peptide synthesis in two different solution systems. In the one-layer system containing water and DMF, Z(20)Csin catalyzed the peptide bond formation in a higher yield than that by unmodifide chymotrypsin and enabled a synthetic reaction in even an 80% (v/v) DMF media, in which the hydrolytic reaction could not be carried out. Z(20)Csin catalyzed the condensation between some N-acyl amino acids or peptide derivatives and amino acids in 90% ethylacetate, 90% hexane or 50% benzene. This latter method employs a two-layer system, and the modified enzyme may be able to reduce the number of synthetic steps when preparing acyl peptides.  相似文献   

11.
Enzymes suspended in organic solvents represent a versatile system for studying the involvement of water in catalytic properties and their flexibility in adapting to different environmental conditions. The extremely halophilic alkaline p-nitrophenylphosphate phosphatase from the archaeon Halobacterium salinarum was solubilized in an organic medium consisting of reversed micelles of hexadecyltrimethylammoniumbromide in cyclohexane, with 1-butanol as cosurfactant. Hydrolysis of p-nitrophenylphosphate was nonlinear with time when the enzyme was microinjected into reversed micelles that contained substrate. These data are consistent with a kinetic model in which the enzyme is irreversibly converted from an initial form to a final stable form during the first seconds of the encapsulation process. The model features a rate constant (k) for that transition and separate hydrolysis rates, v(1) and v(2), for the two forms of the enzyme. The enzyme conversion may be governed by the encapsulation process.  相似文献   

12.
Enhancement effect of water activity on enzymatic synthesis of cephalexin   总被引:3,自引:0,他引:3  
The effect of water activity (a(w)) of the reaction medium on the enzymatic synthesis of cephalexin (CEX) from 7-amino-3-deacetoxycephalosporanic acid (7-ADCA) and D-alpha-phenylglycine methyl ester (PGM) was investigated using the alpha-amino acid ester hydrolase enzyme from Xanthomonas citri. It was found that the synthetic activity of the enzyme and the conversion yield were markedly improved when the a(w) of the reaction medium was lowered to about 0.97. The water activity depressing agents evaluated were glycerol, sucrose, and sorbitol, and the conversion yields were improved up to 170% with 15% glycerol, 230% with 30% sucrose, and 270% with 20% sorbitol, respectively. The extent of favorable effect of a(w) on the conversion yield was not the same among the a(w) depressors, probably due to other unknown interactions between the enzyme and depressors. However, optimal a(w) values corresponding to the maximum conversion yield coincided for all a(w) depressors used. The conversion yield of CEX showed an increasing trend with increasing a(w) up to the optimal a(w) value (0.96-0.97) which corresponds to the maximum conversion yield and a decreasing trend beyond the optimal a(w). There appears to be a delicate balance between the hydrolytic reaction of PGM and synthetic reaction of CEX. The increasing a(w)-[E . PGM] complex and the branched reaction pathway fluxes from [E . PGM] to PG (D-alpha-phenyl glycine) and CEX are balanced in such a way that the maximum CEX conversion yield is obtained at a(w) value of 0.96-0.97. The a(w) depressors stabilized the enzyme somewhat, but this positive effect was considered to be only a minor contribution to the substantial yield enhancement. The a(w) depressor effect on viscosity and in turn the mass transfer rate limitation was ruled out since the change in conversion due to the viscosity change was found to be insignificant. (c) 1993 John Wiley & Sons, Inc.  相似文献   

13.
Steroid bioconversion in a microemulsion system   总被引:1,自引:0,他引:1  
The Delta(1,2)-dehydrogenation of high concentrations of the steroid -methyl-Reichstein's compound S-21-acetate (16MRSA) in a microemulsion system was studied using heat-dried and thawed Arthrobacter simplex cells as biocatalyst. The microemulsion system consists of an organic phase [75-95% (v/v)] with steroid (1-60 g/L(tot)), an aqueous phase [5-25% (v/v)] containing the cells (5-30 g/L(tot)), and a neutral surfactant (5-20 g/L organic solvent). Benzene derivatives, which solubilize 16MRSA up to 94 g/L, and phospholipids were used as organic solvents and surfactants, respectively, and menadione was added as an external electron acceptor. Factors affecting the dehydrogenation rate in the microemulsion system were studied. The influences of the 16MRSA and the menadione concentration on the dehydrogenation rate were described by Michaelis-Menten kinetics, apparent V'(max) and K'(m) values of 2.06 g/g dry weight h and 18.9 g/L for 16MRSA and 4.97 g/g dry weight h and 1.91 g/L for menadione being obtained. Optimal menadione concentration was dependent on the steroid concentration was dependent on the steroid concentration used. The reaction was strongly inhibited by high product concentrations. Much higher activities were obtained with the thawed cells than with the dried cells, conversions of 98% being reached within 14-16 h. for 16MRSA and cell dry weight concentrations of 40 and 10 g/L, respectively. Activity retention in a batch stirred tank reactor remained constant during the first 16-24 h of operation and then decreased, depending on the stirring rate; 22 to 65% of the initial reaction rate was obtained after 48 h at stirring rates of 650 and 2000 rpm, respectively.  相似文献   

14.
A Pseudomonas luteola strain possessing azoreductase activity was utilized to decolorize a reactive azo dye (C. I. Reactive Red 22) with fed-batch processes consisting of an aerobic cell growth stage and an anaerobic fed-batch decolorization stage. The fed-batch decolorization was conducted with different agitation and aeration rates, initial culture volumes, dye loading strategies, and yeast extract to dye (Y/D) ratios, and the effect of those operation parameters on azo dye decolorization was evaluated. Dissolved oxygen strongly inhibited the azo reduction activity; thus aeration should be avoided during decolorization but slight agitation (around 50 rpm) was needed. With the periodical feeding strategy, the specific decolorization rate (v(dye)) and overall decolorization efficiency (eta(dye)) tended to increase with increasing feeding concentrations of dye, whereas substrate inhibition seems to arise when the feeding concentration exceeded 600 mg dye/L. In the continuous feeding mode, higher initial culture volume resulted in better eta(dye) due to higher biomass loading, but lower v(dye) due to lower dye concentration in the bioreactor. With a volumetric flow rate (F) of 25 mL/h, both v(dye) and eta(dye) increased almost linearly with the increase in the loading rate of dye (F(dye)) over the range of 50-200 mg/h, while further increase in F(dye) (400 mg/h) gave rise to a decline in v(dye) and eta(dye). As the F was doubled (50 mL/h), the v(dye) and eta(dye) increased with F(dye) only for F(dye) < 80 mg/h. The best v(dye) (113.7 mg dye g cell(-)(1) h(-)(1)) and eta(dye) (86.3 mg dye L(-)(1) h(-)(1)) were achieved at F(dye) = 200 mg/h and F = 25 mL/h. The yield coefficient representing the relation between dye decolorized and yeast extract consumed was estimated as 0.8 g/g. With F(dye) = 75 mg/h, the Y/D ratio should be higher than 0.5 to ensure sufficient supply of yeast extract for stable fed-batch operations. However, performance of the fed-batch decolorization process was not appreciably improved by raising the Y/D ratio from 0.5 to 1.875 but was more sensitive to the changes in the dye loading rate.  相似文献   

15.
Induction of the bacteriochlorophyll fluorescence under rectangular shape of intense laser diode illumination (1 W cm(-2), 808 nm) was measured over wide time range from 10 mus to 4 s in whole cells, chromatophore and isolated reaction center protein of wild type and carotenoid-less mutant (R-26.1) of purple photosynthetic bacterium Rhodobacter sphaeroides. While the antenna-containing species showed large and positive variable fluorescence (F (v)) to initial fluorescence (F (0)) (F (v)/F (0) approximately 4.5 in whole cells), the isolated RC had negative change (F (v)/F (0) approximately -0.6) during photochemistry. In chromatophore from R-26.1, only seven times higher rate was measured than in isolated reaction center under identical experimental conditions. The enhancement effect of large antenna on the rate of photochemistry in chromatophore was partially compensated by the favorable pigment absorption properties in isolated RC. The transition from membrane bound to isolated form of the reaction center was probed by titration of zwitterionic detergent LDAO in chromatophore, and at 0.03% LDAO concentration, sharp change of the variable fluorescence was observed. The sudden drop was explained by the formation of LDAO micelles. After the photochemical phase, additional change of fluorescence yield could be observed in isolated RC considered as manifestation of long-living conformations of the trapped redox states of the protein characterized by non-exponential kinetics. Strong support was provided for use of the fluorescence induction to track structural and conformation changes at their earliest phases in chromatophores and isolated reaction centers.  相似文献   

16.
Several kinds of modified chymotrypsin were prepared with water-soluble acylating reagents, and their characteristics after hydrolyzing with unmodified chymotrypsin in aqueous-N,N’ -dimethylformamide (DMF) media were compared. It was found that chymotrypsin (Csin), of which a 20% amino group was modified with a benzyloxycarbonyl group (Z(20)Csin), had more favorable characteristics than unmodified chymotrypsin with regard to hydrolytic activity in an aqueous DMF media. We also investigated the Z(20)Csin-catalyzed peptide synthesis in two different solution systems. In the one-layer system containing water and DMF, Z(20)Csin catalyzed the peptide bond formation in a higher yield than that by unmodifide chymotrypsin and enabled a synthetic reaction in even an 80% (v/v) DMF media, in which the hydrolytic reaction could not be carried out. Z(20)Csin catalyzed the condensation between some N-acyl amino acids or peptide derivatives and amino acids in 90% ethylacetate, 90% hexane or 50% benzene. This latter method employs a two-layer system, and the modified enzyme may be able to reduce the number of synthetic steps when preparing acyl peptides.  相似文献   

17.
Kuo CW  Hung HC  Tong L  Chang GG 《Proteins》2004,54(3):404-411
Human mitochondrial NAD(P)+-dependent malic enzyme was strongly inhibited by Lu3+. The X-ray crystal structures indicated a structural change between the metal-free and Lu3+-containing enzymes (Yang Z, Batra R, Floyd DL, Hung HC, Chang GG, Tong L. Biochem Biophys Res Commun 2000;274:440-444). We characterized the reversible slow-binding mechanism and the structural interconversion between Mn2+- and Lu3+-containing human mitochondrial malic enzymes. When Lu3+ was added, the activity of the human enzyme showed a downward curve over time, similar to that of the pigeon enzyme. The rate of the transformation (k(obs)) from the initial rate to the steady-state rate increased hyperbolically with the concentration of Lu3+, suggesting the involvement of an isomerization step. Lu3+ had a much higher affinity for the isomerized form (K*(i,Lu (app)) = 4.8 microM) than that of the native form (K(i,Lu (app)) = 148 microM). When an excess of Mn2+ was added to the Lu3+-inhibited enzyme, assays of the kinetic activity showed an upward trend, indicating reactivation. This result also indicated that the reactivation was a slow process. Fluorescence quenching experiments confirmed that the Lu3+-induced isomerization was completely reversible. The dynamic quenching constants for the metal-free, Mn2+-containing, and Lu3+-containing enzyme were 3.08, 3.07, and 3.8 M(-1), respectively. When the Lu3+-containing enzyme was treated with excess Mn2+, the dynamic quenching constant returned to the original value (3.09 M(-1)). These results indicated that binding of Mn2+ did not induce any conformational change in the enzyme. The open form transformed to the closed form only after substrate binding. Lu3+, on the other hand, transformed the open form into a catalytically inactive form. Excess Mn2+ could replace Lu3+ in the metal binding site and convert the inactive form back into the open form. This reversible process was slow in both directions because of the same but opposite structural change involved.  相似文献   

18.
The reaction rates (k(r)) of 5,7-diisopropyl-tocopheroxyl radical (Toc) with catechins (epicatechin (EC), epicatechin gallate (ECG), epigallocatechin (EGC), epigallocatechin gallate (EGCG)) and related compounds (methyl gallate (MG), 4-methylcatechol (MC), and 5-methoxyresorcinol (MR)) have been measured by stopped-flow spectrophotometer. The k(r) values increased in the order of MR < < MG < EC < MC approximately ECG < EGC < EGCG in ethanol and 2-propanol/H(2)O (5/1, v/v) solutions, indicating that the reactivity of the OH groups in catechins increased in the order of resorcinol A-ring < < gallate G-ring < catechol B-ring < pyrogallol B-ring. The catechins which have lower oxidation potentials show higher reactivities. The rate constants for catechins in micellar solution showed notable pH dependence with one or two peaks around pH 9-11, because of the dissociation of various phenolic hydroxyl protons in catechins. The structure-activity relationship in the free-radical-scavenging reaction by catechins has been clarified by the detailed analyses of the pH dependence of k(r) values. The reaction rates increased remarkably with increasing the anionic character of catechins, that is, the electron-donating capacity of catechins. The mono anion form at catechol B-and resorcinol A-rings and dianion form at pyrogallol B-and gallate G-rings show the highest activity for free-radical-scavenging. It was found that catechins (EC, ECG, EGC, and EGCG) have activity similar to or higher than that of vitamin C in vitamin E regeneration at pH 7-12 in micellar solution.  相似文献   

19.
The effect of glycerol on the hydrolytic activity of thermolysin (EC 3.4.24.4) has been compared with the effect on the condensation of N-benzyloxycarbonyl-L-aspartic acid with L-phenylalanine methyl ester to form N-benzyloxycarbonyl-L-aspartyl-L-phenylalanine methyl ester (Z X Asp X Phe X OMe), the precursor to the sweet-tasting compound L-aspartyl-L-phenylalanine methyl ester. Hydrolytic activity was measured by the degradation of azocasein and furylacryloyl-L-glycyl-L-leucinamide. Increasing concentrations of glycerol reversibly inhibited the hydrolytic activity of the enzyme toward both substrates. The inclusion of glycerol in the synthetic medium facilitated the production of Z X Asp X Phe X OMe in a water-soluble system but reduced the initial rate of peptide synthesis. Glycerol stabilized thermolysin against thermal denaturation.  相似文献   

20.
Feasibility of elemental sulfur reduction by Desulfovibrio desulfuricans in anaerobic conditions in a stirred reactor was studied. Hydrogen was used as energy source, whereas the carbonated species were bicarbonate and yeast extract. Attention was paid to reactor engineering aspects, biofilm formation on the sulfur surface, hydrogen sulfide formation rate and kinetics limitations of the sulfur reduction. D. desulfuricans formed stable biofilms on the sulfur surface. It was found that active sulfur surface availability limits the reaction rate. The reaction rate was first order with respect to sulfur and hydrogen velocity had no effect in the reaction rate for the range 8.2 x 10(-2) to 4.1 x 10(-1) Nm(3) m(-2) min(-1). At a superficial gas velocity (u(G)) = 3.1 x 10(-2) Nm(3) m(-2) min(-1), H(2)S(g) production rate decreased due to a deficient H(2)S stripping. A maximum H(2)S(g) production rate of 2.1 g H(2)S L(-1) d(-1) was achieved during 5 days with an initial sulfur density of 4.7% (w/v).  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