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 187 毫秒
1.
In this study, a biofilm reactor with plastic composite support (PCS), made by high-temperature extrusion of agricultural products and polypropylene, was evaluated for nisin production using L. lactis strain NIZO 22186. The high-biomass density of the biofilm reactor was found to contribute to a significantly shorter lag time of nisin production relative to a suspended-cell reactor. In comparison to glucose (579 IU/mL), sucrose significantly increased the nisin production rate by 1.4-fold (1100 IU/mL). However, results revealed that high levels of sucrose (8% w/v) had a suppressing effect on nisin production and a stimulating effect on lactic acid production. A high concentration of MgSO4.7H2O at 0.04% (w/v) was found to reduce the nisin production, while concentrations of KH2PO4 of up to 3% (w/v) did not have any significant effect on growth or nisin production. The best of the tested complex media for nisin production using the PCS biofilm reactor consisted of 4% (w/v) sucrose, 0.02% (w/v) MgSO4.7H2O, and 0.1% (w/v) KH2PO4. Nisin production rate in the biofilm reactor was significantly increased by 3.8-fold (2208 IU/mL) when using the best complex medium tested.  相似文献   

2.
Omeprazole was found to inhibit the K+-stimulated ATPase activity of the gastric (H+ + K+)-ATPase in parallel with the K+-stimulated p-nitrophenylphosphatase activity and the phosphoenzyme formation. The degree of inhibition of ATPase activity was directly correlated to the amount inhibitor bound to the enzyme preparation down to about 15% of the control enzyme activity. The acid-decomposed form of omeprazole, i.e. the inhibitory form, was found to react with and bind to sulfhydryl groups within the (H+ + K+)-ATPase preparation with close to a 1:1 stoichiometry. beta-Mercaptoethanol, when added beforehand and in a 10-fold excess of omeprazole, completely prevented binding of the inhibitor and its inhibition of the enzyme. In the presence of beta-mercaptoethanol two different reaction products could be detected in addition to omeprazole; the reduced form of omeprazole (H 168/22), and a product formed between beta-mercaptoethanol and a decomposition product, generated from omeprazole. Under those conditions neither inhibition nor binding was obtained, indicating that none of these three compounds was the inhibitor. Rather, the compound generated from omeprazole and reacting rapidly with either beta-mercaptoethanol or the -SH groups of the enzyme was the likely inhibitor compound. In order to reverse already established inhibition higher concentrations of beta-mercaptoethanol were needed than for protection indicating two different reaction pathways for protection and reversal by beta-mercaptoethanol. The reversal reaction was explained by a two-step reaction; in the first step the bound inhibitor was exchanged for a beta-mercaptoethanol molecule resulting in formation of compound H 168/22 and a mixed disulfide between the enzyme and beta-mercaptoethanol. In the second step, attack of another beta-mercaptoethanol molecule results in liberation of active enzyme and generation of the disulfide form of beta-mercaptoethanol. This hypothesis was substantiated by the fact that when 1 mM beta-mercaptoethanol was added to inhibited enzyme the radiolabel was partially displaced, without any change in the concentration of modified -SH groups.  相似文献   

3.
The interaction was studied between the mitochondrial enzyme thiosulfate sulfurtransferase and liposomes, in the form of large unilamellar vesicles (LUV), prepared from either cardiolipin (CL), PtdCho or PtdSer. At equivalent concentrations of lipid, more partially folded thiosulfate sulfurtransferase bound to CL/LUV than to PtdSer/LUV, and only traces were bound to PtdCho/LUV. Native thiosulfate sulfurtransferase did not bind to any of these LUV. We show that CL/LUV-sequestered thiosulfate sulfurtransferase is inactive but may be reactivated (approximately 56%) with the aid of detergents, thiosulfate, beta-mercaptoethanol and phosphate buffer. Reactivations in the presence of PtdSer/LUV or PtdCho/LUV was only 9% or 1%, respectively. Analysis of the complex by protease digestion and fluorescence spectroscopy indicated that thiosulfate sulfurtransferase was held by CL/LUV and PtdSer/LUV as a folding intermediate. Data presented here suggest that detergents may not interact directly with the protein, but, rather, their primary role in reactivation is to disrupt the LUV, allowing flexibility to the anchored thiosulfate sulfurtransferase molecule, thereby promoting folding. These studies complement other reports which imply a possible role for CL in protein translocation across the mitochondria, since we find that CL binds to thiosulfate sulfurtransferase and sequesters it in a translocation-competent prefolded conformation, which may readily lead to a correctly folded enzyme.  相似文献   

4.
Trehalose and glycerol are known as good stabilizers of function and structure of several macromolecules against stress conditions. We previously reported that they have comparable effectiveness on protecting two yeast cytosolic enzymes against thermal inactivation. However, enzyme protection has always been associated to a decrease in catalytic activity at the stabilizing conditions i.e., the presence of the protective molecule. In the present study we tested trehalose and glycerol on thermal protection of the mammalian cytosolic enzyme phosphofructokinase. Here we found that trehalose was able to protect phosphofructokinase against thermal inactivation as well as to promote an activation of its catalytic activity. The enzyme incubated in the presence of 1 M trehalose did not present any significant inactivation within 2 h of incubation at 50 degrees C, contrasting to control experiments where the enzyme was fully inactivated during the same period exhibiting a t0.5 for thermal inactivation of 56+/-5 min. On the other hand, enzyme incubated in the presence of 37.5% (v/v) glycerol was not protected against incubation at 50 degrees C. Indeed, when phosphofructokinase was incubated for 45 min at 50 degrees C in the presence of lower concentrations of glycerol (7.5-25%, v/v), the remaining activity was 2-4 times lower than control. These data show that the compatibility of effects previously shown for trehalose and glycerol with some yeast cytosolic enzymes can not be extended to all globular enzyme system. In the case of phosphofructokinase, we believe that its property of shifting between several different complex oligomers configurations can be influenced by the physicochemical properties of the stabilizing molecules.  相似文献   

5.
Prostaglandin (PG) E synthase was solubilized with 6 mM sodium deoxycholate from the microsomal fraction of bovine hearts. The enzyme was purified by about 800-fold to apparent homogeneity. The specific activity of the purified enzyme was about 830 mU/mg of protein, and the K(m) value for PGH(2) was 24 microM. The molecular weight of the enzyme was about 31000 on SDS-polyacrylamide gel electrophoresis and was about 60000 by gel filtration. The enzyme was separated from glutathione (GSH) S-transferase by DEAE-Toyopearl column chromatography, and did not exhibit any GSH S-transferase activity towards four different substrates. The purified enzyme was active in the absence of GSH, but it was activated by various SH-reducing reagents including dithiothreitol, GSH, or beta-mercaptoethanol. This is the first reported purification of membrane-bound PGE synthase to apparent homogeneity.  相似文献   

6.
Formation of noncovalent complexes between alpha-chymotrypsin (CT) and a polyelectrolyte, polybrene (PB), has been shown to produce two major effects on enzymatic reactions in binary mixtures of polar organic cosolvents with water. (i) At moderate concentrations of organic cosolvents (10% to 30% v/v), enzymatic activity of CT is higher than in aqueous solutions, and this activation effect is more significant for CT in complex with PB (5- to 7-fold) than for free enzyme (1.5- to 2.5-fold). (ii) The range of cosolvent concentrations that the enzyme tolerates without complete loss of catalytic activity is much broader. For enhancement of enzyme stability in the complex with the polycation, the number of negatively charged groups in the protein has been artificially increased by using chemical modification with pyromellitic and succinic anhydrides. Additional activation effect at moderate concentrations of ethanol and enhanced resistance of the enzyme toward inactivation at high concentrations of the organic solvent have been observed for the modified preparations of CT in the complex with PB as compared with an analogous complex of the native enzyme. Structural changes behind alterations in enzyme activity in water-ethanol mixtures have been studied by the method of circular dichroism (CD). Protein conformation of all CT preparations has not changed significantly up to 30% v/v of ethanol where activation effects in enzymatic catalysis were most pronounced. At higher concentrations of ethanol, structural changes in the protein have been observed for different forms of CT that were well correlated with a decrease in enzymatic activity. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 55: 267-277, 1997.  相似文献   

7.
Unlike other vertebrate mevalonate kinase, the enzyme partially purified from neonatal chick liver was not activated by the -SH group protectors reduced glutathione, cysteine, dithiothreitol and beta-mercaptoethanol at any concentrations assayed (0.01-10.00 mM). However, the activity was found to be sensitive to thiol group binding reagents. p-Hydroxymercuribenzoate was the most active inhibitor. At 0.1 mM concentration, p-HMB completely abolished the enzyme activity. N-ethylmaleimide (0.01-1.00 mM) was practically ineffective. Inhibition by p-HMB was temperature dependent, being more potent at 37 degrees C than at 4 degrees C.  相似文献   

8.
A basic serine protease which is active on casein and fibrinogen was purified from Bothrops moojeni venom using a single step chromatography on a CM-Sepharose fast flow column. The enzyme, MOO3, was not hemorrhagic and presented only a trace of blood-clotting activity. Synthetic chromogenic substrates (azoacasein and azoalbumin) where not hydrolyzed by MOO3. Using polyacrylamide gel electrophoresis at pH 4.3, MOO3 showed as a single protein band. Using sodium dodecyl sulfate-polyacrylamide electrophoresis, MOO3 behaved as a single-chain protein with an approximate mol. weight of 27,000, both in the presence and absence of beta-mercaptoethanol. Its pI was 7.8 by electrofocusing. The enzyme did not contain neutral carbohydrates and its N-terminal amino acid was alanine. The amino acid composition showed 249 residues/mole, a high content of hydrophilic amino acids and 14 half-cystine residues, which should account for 7 disulfide bonds. The protease cleaved the A-alpha chain faster than the B-beta of bovine fibrinogen and showed no effect on the delta-chain. Specific esterolytic activity of MOO3 on alpha-N-tosyl-l-arginine methyl ester was 29.64 mumol min-1 x mg-1. MOO3 represented 1.42% (w/w) of the initial desiccated venom. Its proteolytic activity was inhibited by beta-mercaptoethanol, leupeptin, phenylmethylsulphonyl fluoride and ethylenediamine tetraacetate.  相似文献   

9.
The first ribonuclease (RNase) from the Cytophaga-Flavobacterium-Bacteroides phylum, dominant in the marine environment, and also from the first Bizionia species isolated from the tropics was purified and characterized. Extracellular RNase production occurred when the culture medium contained 5-7% (w/v) NaCl. The 53.0 kDa enzyme was purified 29 folds with a recovery of 4% and specific activity of 630unit/mg protein. The pH and temperature optima are 6.5 and 35 degrees C, respectively and the enzyme retains more than half of its activity (relative to optimal assay conditions) after 1h pre-incubation separately with 5% (w/v) NaCl or from pH 5.0 to 8.5 or at 50 degrees C. Dithiothreitol and beta-mercaptoethanol do not inhibit whereas human placental RNase inhibitor protein halves the RNase activity. While Mg(2+), Ba(2+) and Ca(2+) enhanced the enzyme activity, Fe(2+), Cu(2+) and Hg(2+) inactivated it. This RNase degrades uracil containing nucleic acids only. Our isolate could be a novel renewable source of deoxyribonuclease (DNase)--free RNase enzyme.  相似文献   

10.
The effects of important reaction parameters for enhancing isoamyl acetate formation through lipase-catalyzed esterification of isoamyl alcohol were investigated in this study. Increase in substrate (acid) concentration led to decrease in conversions. A critical enzyme concentration of 3 g l(-1) was detected for a substrate concentration of 0.06 M (each of alcohol and acid). Solvents with partition coefficient higher than 1000 (log P>3.0) supported enzyme activity to give high conversions. Acetic acid at higher concentrations could not be esterified easily probably owing to its role in lowering the microaqueous pH of the enzyme. Extraneous water/buffer addition decreased the isoamyl acetate yields slightly ( approximately 10%) at 0.005-0.01% v/v of the reaction mixture and drastically (>40%) at above 0.01% v/v. Buffer saturation of the organic solvent employed improved esterification (upto two-fold), particularly at moderately higher substrate concentrations (>0.18 M). Employing acetic anhydride instead of acetic acid resulted in a two-fold increase in the yields (at 0.25 M substrate). Use of excess nucleophile (alcohol) concentration by increasing the alcohol/acid molar ratio resulted in higher conversions in shorter duration (upto eight-fold even at 1.5 M acetic acid). Yields above 80% were achieved with substrate concentrations as high as 1.5 M and more than 150 g l(-1) isoamyl acetate concentrations were obtained employing a relatively low enzyme concentration of 10 g l(-1). The operational stability of lipase was also observed to be reasonably high enabling ten reuses of the biocatalyst.  相似文献   

11.
Abstract Intracellular acidification has been considered one of a number of mechanisms underlying the inhibition of growth and fermentation by ethanol in yeast. However, most of the studies on the effect of ethanol on yeast intracellular pH (pHi) were carried out by using unadapted cells to which ethanol was added. In this paper we show that the pHi of exponential cells of Saccharomyces cerevisiae IGC 3507 III grown in a medium with glucose and inhibitory concentrations of ethanol only decreased to values below those in unstressed cells (6.9) for concentrations equal to or above 7% (v/v). Only at these supracritical levels (7–10% (v/v)) was pH homeostasis in ethanol-adapted yeast affected. This is consistent with the significant increase of plasma membrane permeability and decrease of plasma membrane H+-ATPase in comparison with the corresponding values in unstressed cells. These deleterious effects were only observed with those high concentrations of toxin. These results indicate that intracellular acidification does not account for inhibition of yeast growth in the presence of ethanol. In fact, growth was inhibited by ethanol concentrations (3–6% (v/v)) that did not lead to the decrease of pHi. Furthermore, even for supracritical concentrations, close to the maximal that allowed growth (10% (v/v)), the dedrease of pHi was not important reaching, at the most, values of 6.5–6.6.  相似文献   

12.
An apparently latent phosphatase which migrated as a protein of Mr 130,000 during sucrose density centrifugation, and a spontaneously active phosphatase (Mr 68,000) were isolated from bovine aortic smooth muscle. Basal phosphorylase phosphatase activity of the latent preparations was stimulated 12 fold by low concentrations of lysine-rich histone-H1 (30 micrograms/ml) and 6 fold by polylysine (Mr 17,000; 12 micrograms/ml), whereas the spontaneously active enzyme was only slightly affected. The enzymatic activity of the spontaneously active preparation was completely destroyed by beta-mercaptoethanol. In contrast, the apparently latent enzyme was converted to a more active form of lower molecular weight (Mr 86,000) following treatment with beta-mercaptoethanol and this form of the enzyme was still stimulateable by histone-H1. These findings show that the aortic spontaneous and apparently latent phosphatase actives are ascribable to separate enzymes and they suggest that the activity of latent phosphatase in living cells may be modulated by cationic proteins such as histones or similar effector molecules.  相似文献   

13.
Summary Addition of hemoglobin, methemoglobin, hemin or hematin in the assay mixture of rat liver 3-hydroxy-3-methylglutaryl CoA (HMGCoA) reductase inhibited the activity of the enzyme. The inhibition by hemin was rapid, without any apparent dependence on time of preincubation. At 20 M hemin, a maximum of about 50% inhibition was obtained in the case of the microsomal enzyme while the solubilized enzyme showed almost 80%6 inhibition. Dithiothreitol at high concentrations or either of the two substrates of the enzyme (HMGCoA and NADPH) could afford partial protection when added before hemin. The Km for both the substrates increased in the presence of hemin. The inhibition by hemin appeared to be irreversible, the presence of KCN or NaN3 being the only means of preventing the inhibition. Molecular oxygen was required for the inhibition. Oxygen radicals and H2O2, however, did not seem to be involved. This offered a clue that an oxidation reaction of the reductase protein may be the likely mechanism of its inactivation. The enzyme protein did not, however, get degraded under the conditions of inhibition.Abbreviations HMGCoA 3-Hydroxy-3-methylglutaryl coenzyme - DTT Dithiothreitol - DTNB 5,5-Dithiobis-(2-nitrobenzoic acid) - SDS-PAGE Sodium dodecyl sulphate-polyacrylamide gel electrophoresis  相似文献   

14.
The major challenge in developing cryopreservation protocols for microencapsulated cells is that the relatively large size (300-400 microm) and the fragile semipermeable membrane of microcapsules makes them particularly prone to cryodamage. Rapid-cooling cryopreservation protocols with high DMSO concentrations (3.5M, 25% v/v) resulted in low post-thaw cell viability (<10%), which did not improve with higher concentrations (4.5M, 32% v/v) and longer exposure to DMSO, even though the majority of microcapsules (60-80%) remained intact. Subsequent investigations of slow cooling with a range of DMSO and EG concentrations resulted in a much higher post-thaw cell viability (80-85%), with the majority of the microcapsules remaining intact ( approximately 60%) when DMSO was used at a concentration of 2.8M (20% v/v) and EG at a concentration of 2.7M (15% v/v). The presence of 0.25M sucrose significantly improved post-thaw cell viability upon slow cooling with 2.8M (20% v/v) DMSO, although it had no effect on microcapsule integrity. Multistep exposure and removal of sucrose did not significantly improve either post-thaw cell viability or microcapsule integrity, compared to a single-step protocol. Ficoll 20% (w/v) also did not significantly improve post-thaw cell viability and microcapsule integrity. Hence, the optimal condition for microcapsule cryopreservation developed in this study is slow cooling with 2.8M (20% v/v) DMSO and 0.25M sucrose.  相似文献   

15.
Morphine (200 micrograms/rat) was injected intraventricularly (i.v.t.) into normal and into long-term castrated (4 weeks) adult male rats. Animals were killed 10, 20, 40 and 60 min after treatment. In normal animals, the treatment with morphine resulted in a significant increase of serum prolactin concentrations at all time intervals considered. However, the i.v.t. injection of 200 micrograms morphine/rat into castrated rats did not exert any significant effect on prolactin release at any time interval considered. When morphine (200 micrograms/rat) was administered i.v.t. together with the specific opioid receptor blocker naloxone (7.5 or 15 micrograms/rat) the stimulatory effect of morphine on prolactin release was diminished at 10 min, and totally blocked at 20 min. Naloxone given alone did not influence serum prolactin concentrations. The results suggest that the presence of endogenous androgens is essential to permit the stimulatory effect of morphine on prolactin release.  相似文献   

16.
A high molecular weight phosphoprotein phosphatase was purified from rabbit liver using high speed centrifugation, acid precipitation, ammonium sulfate fractionation, chromatography on DEAE-cellulose, Sepharose-histone, and Bio-Gel A-0.5m. The purified enzyme showed a single band on a nondenaturing polyacrylamide anionic disc gel which was associated with the enzyme activity. The enzyme was made up of equimolar concentrations of two subunits whose molecular weights were 58,000 (range 58,000-62,000) and 35,000 (range 35,000-38,000). Two other polypeptides (Mr 76,000 and 27,000) were also closely associated with our enzyme preparation, but their roles, if any, in phosphatase activity are not known. The optimum pH for the reaction was 7.5-8.0. Km value of phosphoprotein phosphatase for phosphorylase a was 0.10-0.12 mg/ml. Freezing and thawing of the enzyme in the presence of 0.2 M beta-mercaptoethanol caused an activation (100-140%) of phosphatase activity with a concomitant partial dissociation of the enzyme into a Mr 35,000 catalytic subunit. Divalent cations (Mg2+, Mn2+, and Co2+) and EDTA were inhibitory at concentrations higher than 1 mM. Spermine and spermidine were also found to be inhibitory at 1 mM concentrations. The enzyme was inhibited by nucleotides (ATP, ADP, AMP), PPi, Pi, and NaF; the degree of inhibition was different with each compound and was dependent on their concentrations employed in the assay. Among various types of histones examined, maximum activation of phosphoprotein phosphatase activity was observed with type III and type V histone (Sigma). Further studies with type III histone indicated that it increased both the Km for phosphorylase a and the Vmax of the dephosphorylation reaction. Purified liver phosphatase, in addition to the dephosphorylation of phosphorylase a, also catalyzed the dephosphorylation of 32P-labeled phosphorylase kinase, myosin light chain, myosin, histone III-S, and myelin basic protein. The effects of Mn2+, KCl, and histone III-S on phosphatase activity were variable depending on the substrate used.  相似文献   

17.
Increased production, secretion, and activity of β-glucosidase in the filamentous fungus Termitomyces clypeatus was achieved in presence of the glycosylation inhibitor 2-deoxy-d-glucose (0.05%, w/v) during submerged fermentation. Enzyme activity increased to 163 U/mL by adding mannose (2 mg/mL) to the medium. Such a high enzyme activity has not been achieved without mutation or genetic manipulation. The Km and Vmax of the enzyme in culture medium were determined to be 0.092 mM and 35.54 U/mg, respectively, with p-nitrophenyl β-d-glucopyranoside as substrate, confirming its high catalytic activity. The enzyme displayed optimum activity at pH 5.4 and 45°C. The enzyme was fairly stable between acidic to alkaline pH and retained about 75 ∼ 65% residual activities between pH 4 and 10.6 and demonstrated full activity at 45°C for 3 days. The enzyme was also stable in the presence of Zn2+ and Mg2+ and 80% of the residual activity was observed in the presence of Mn2+, Ca2+, K+, Cu2+, EDTA, and sodium azide. Around 70% of the activity was retained in the presence of 2 M guanidium HCl and 3 M urea, whereas the activity was 5 and 2 times higher in the presence of 4 mM beta-mercaptoethanol and 50 mM DTT, respectively. The enzyme obtained from the culture filtrate showed potential cellulose saccharifying ability which increased further when supplemented with commercial cellulase. Thus, this enzyme could be used without any additional downstream processing for commercial cellulase preparation and production of bioethanol or for other biotechnological applications.  相似文献   

18.
Papain kinetics in the presence of a water-miscible organic solvent   总被引:5,自引:0,他引:5  
The effects of various concentrations of a water-miscible organic solvent [a 7:3 (v/v) mixture of N, N dimethylformamide and dimethylsulfoxide] on the kinetics of papain have been investigated. The parameters k(cat) and K(m) for the amidase and esterase activity of papain using N-alpha-benzoyl-DL-arginine-p-nitroanilide (BAPNA) and N-alpha-benzoyl-L-arginine ethyl ester (BAEE) as substrates were determined. For both types of activity, k(cat) initially increased (up to about 15% solvent), and then decreased with increasing concentrations of organic solvent. In contrast, K(m) increased sharply with the organic solvent concentration. Active site titration at 0 and 50% solvent indicated no change in the amount of active enzyme. Fluorometric measurements of the emission spectrum of papain did not indicate any major conformational changes with increasing concentrations of organic solvent.  相似文献   

19.
Changes in unfolding and enzymatic activity of bovine carbonic anhydrase II (BCA II) in different concentrations of 2,2,2-trifluoroethanol (TFE) were investigated by 1-anilino-8-naphthalenesulfonate (ANS) fluorescence emission spectra, far-UV CD spectra, and enzyme activity. The results showed that the activity and conformation of BCA II changed according to the concentration of TFE. Significant aggregation was observed when BCA II was denatured at TFE concentrations between 10 and 35% (v/v). When the concentration of TFE exceeded 40%, the aggregation of BCA II was not very obvious. The activity of BCA II decreased almost to zero as the TFE concentration reached 26%. The ANS fluorescence spectra indicated the tertiary conformations of BCA II were more stable in solutions with TFE concentrations lower than 15% (v/v) and higher than 40% (v/v). Far-UV CD spectra showed that high concentrations (higher than 25%) of TFE could induce BCA II to form more alpha-helix structures and caused these structures to be in relatively stable states. The native conformation of BCA II being destroyed after its inactivity indicated that the active sites of BCA II is situated in a limited region and has more flexibility than the whole enzyme molecule.  相似文献   

20.
Changes in unfolding and enzymatic activity of bovine carbonic anhydrase II (BCA II) in different concentrations of 2,2,2-trifluoroethanol (TFE) were investigated by 1-anilino-8-naphthalenesulfonate (ANS) fluorescence emission spectra, far-UV CD spectra, and enzyme activity. The results showed that the activity and conformation of BCA II changed according to the concentration of TFE. Significant aggregation was observed when BCA II was denatured at TFE concentrations between 10 and 35% (v/v). When the concentration of TFE exceeded 40%, the aggregation of BCA II was not very obvious. The activity of BCA II decreased almost to zero as the TFE concentration reached 26%. The ANS fluorescence spectra indicated the tertiary conformations of BCA II were more stable in solutions with TFE concentrations lower than 15% (v/v) and higher than 40% (v/v). Far-UV CD spectra showed that high concentrations (higher than 25%) of TFE could induce BCA II to form more α-helix structures and caused these structures to be in relatively stable states. The native conformation of BCA II being destroyed after its inactivation indicated that the active site of BCA II is situated in a limited region and has more flexibility than the whole enzyme molecule.  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