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 0 毫秒
1.
4-Hydroxy-2-keto-pentanoic acid aldolase from Escherichia coli was identified as a class I aldolase. The enzyme was found to be highly selective for the acetaldehyde acceptor but would accept α-ketobutyric acid or phenylpyruvic acid in place of the pyruvic acid carbonyl donor.Aldolase-catalyzed reactions are involved in the latter stages of many bacterial catabolic pathways responsible for the degradation of aromatic compounds (3). In particular, many of the bacterial meta cleavage pathways for aromatic degradation proceed via the common intermediate 4-hydroxy-2-keto-pentanoic acid (HKP) (3), whose cleavage to acetaldehyde and pyruvic acid (Fig. (Fig.1)1) is catalyzed by an aldolase enzyme. Previous biochemical work on HKP aldolase is limited to the enzyme activity from Pseudomonas strains. HKP aldolase from Pseudomonas sp. strain CF600 was shown to be activated by Mn2+ ions (9), whereas 4-hydroxy-4-methyl-2-oxoglutarate aldolase from Pseudomonas putida (11) requires Mg2+ ions for catalytic activity, and these enzymes appear to fall into the class II family of aldolases. Open in a separate windowFIG. 1Reaction catalyzed by HKP aldolase. Also illustrated are the previous reaction on the phenylpropionate catabolic pathway, the lactone derivative of HKP, and the method used for the coupled enzyme assay.It has been reported previously that the HKP aldolase activity in extracts of Escherichia coli would process both enantiomers of the substrate, unlike the corresponding activities from Pseudomonas and Acinetobacter spp., which were enantioselective (2). We have previously established that the preceding enzyme on the phenylpropionic acid catabolic pathway, namely, 2-hydroxypentadienoic acid hydratase (MhpD), catalyzes a stereospecific hydration reaction (8). We therefore wished to examine the stereoselectivity and substrate selectivity of HKP aldolase from E. coli.HKP aldolase activity was detectable in extracts of E. coli W3110 in a stopped assay involving treatment of HKP with extract for 30 min, followed by heat treatment and then addition of lactate dehydrogenase and NADH. Levels of enzyme activity were low (7.0 mU/mg of protein) and were not enhanced by inclusion of phenylpropionic acid in the growth media (1). The enzyme was purified by precipitation with 50 to 80% ammonium sulfate, phenyl-agarose hydrophobic-interaction chromatography, Q Sepharose anion-exchange fast protein liquid chromatography (FPLC), and Mono Q anion-exchange FPLC. The purified enzyme had a specific activity of 184 mU/mg of protein, a 26-fold purification overall. Although not purified to homogeneity, the purified enzyme was entirely free of background NADH oxidase activity and could be examined in a continuous assay by incubation with lactate dehydrogenase and NADH. Maximum activity was obtained in the pH range 6.25 to 6.75, with sharp inflections of activity at pH 6.0 and 8.0.The purified HKP aldolase showed no observable dependence on divalent metal ions, unlike the purified Pseudomonas aldolases (9, 11). Furthermore, treatment of the enzyme with the metal chelator EDTA at 10 mM resulted in no loss of enzyme activity. Moreover, treatment of enzyme with sodium borohydride in the presence of substrate HKP resulted in 100% loss of activity, indicative of an imine linkage. Treatment with sodium borohydride in the absence of substrate resulted in only a slight (<10%) loss of activity; thus, the imine linkage is formed only upon addition of the substrate. These data imply that the E. coli enzyme is a class I aldolase utilizing an imine linkage between the C-2 carbonyl of HKP and the ɛ-amino group of a lysine residue at the active site. Since 80% amino acid sequence identity has been determined between the Pseudomonas strain CF600 DmpG and E. coli MhpE gene products corresponding to the respective HKP aldolases (4), the difference in behavior between the E. coli and Pseudomonas enzymes is most surprising. We note that although the Pseudomonas HKP aldolase was reported to show a six- to eightfold activation by Mn2+, the enzyme retained residual activity after EDTA treatment and the presence of an imine linkage was not investigated (9).The stereoselectivity of the enzymatic reaction was examined by treatment of racemic HKP with the enzyme for various reaction times, followed by acid-catalyzed lactonization of the remaining HKP substrate to give 2-keto-4-methyl-γ-butyrolactone, followed by organic acids HPLC analysis. Time-dependent consumption of HKP was observed, but after long reaction times approximately 60% of the substrate remained (Fig. (Fig.2A).2A). Kinetic studies subsequently showed that the equilibrium position lies strongly in favor of the forward reaction; therefore, these data imply that the enzyme utilizes only one enantiomer of the substrate. Since the earlier study of the E. coli enzyme stereospecificity was carried out with crude extract (2), it is possible that there is a second aldolase with the opposite stereospecificity in E. coli, although we observed only one peak of activity upon enzyme purification. Open in a separate windowFIG. 2Analysis of HKP aldolase-catalyzed reaction via HPLC analysis of the lactone derivative α-methyl-γ-methyl-γ-butyrolactone. (A) Forward reaction using racemic HKP as the substrate (percentage of initial peak area). (B) Reverse reaction using pyruvic acid and acetaldehyde as substrates (concentrations determined by peak area, versus authentic standards).The reverse reaction was assayed by incubation of enzyme with acetaldehyde and pyruvate, followed by lactonization of the HKP product under acidic conditions, and HPLC analysis. Time-dependent formation of product was observed (Fig. (Fig.2B),2B), indicating that HKP aldolase also catalyzes the reverse reaction. Concentrations of product were deduced by calibration with known amounts of synthetic 2-keto-4-methyl-γ-butyrolactone (10); thus, after 24 h the reaction mixture contained 3.7 mM HKP. Initial concentrations of acetaldehyde and pyruvic acid were 360 mM and 180 mM, respectively; thus, a Keq of 17 M in favor of the forward reaction can be deduced. From the rate of reaction over the first hour, it was calculated that the reverse reaction proceeds at 13% of the rate of the forward reaction.The substrate selectivity for the reverse reaction catalyzed by HKP aldolase was examined with respect to the carbonyl donor pyruvic acid and the carbonyl acceptor acetaldehyde. No product formation was observed when propionaldehyde was used in place of acetaldehyde; thus, the enzyme is highly selective for the carbonyl acceptor. However, lactone products with similar retention times and λmaxs were observed by HPLC using either α-ketobutyric acid or phenylpyruvic acid as the substrate for the reverse reaction. The apparent rates of formation of the new lactone derivatives are similar to those observed with pyruvic acid as the substrate (Table (Table1);1); thus, α-ketobutyric acid and phenylpyruvic acid appear to be converted efficiently by the enzyme.

TABLE 1

Substrate selectivity for the MhpE-catalyzed forward and reverse reactionsa
Substrate(s) (concn)Assay method (retention time)bvrelc
4-Hydroxy-2-keto-pentanoic acid (200 μM)LDH/NADH1.00
Acetaldehyde + pyruvic acid (both 100 mg/ml)HPLC (34 min)0.13
Propionaldehyde + pyruvic acid (both 100 mg/ml)HPLC
Acetaldehyde + 2-keto-butyric acid (both 100 mg/ml)HPLC (37 min)0.12
Acetaldehyde + phenylpyruvic acid (both 100 mg/ml)HPLC (33 min)0.13
Open in a separate windowaAssay methods for forward and reverse reactions are described in the text and are illustrated in Fig. Fig.1.1. Assays were conducted at 20°C in 50 mM potassium phosphate buffer, pH 7.0. bRetention time for α-keto-γ-butyrolactone derivatives observed on organic acids HPLC (Bio-Rad HPX-87H column, eluent 0.005 mM H2SO4, flow rate 0.6 ml/min). LDH, lactate dehydrogenase. cReaction rate relative to HKP forward reaction. —, no reaction observed. In summary, HKP aldolase from E. coli is shown to be a class I aldolase enzyme, which proceeds via an imine linkage between the substrate and the enzyme active site. Only a small number of class I aldolases have been found in prokaryotes (57), including an aldolase enzyme involved in bacterial naphthalene sulfonate degradation (6). The partially purified enzyme is shown to be selective for one enantiomer of the substrate, presumably the 4S enantiomer produced by the preceding enzyme on the pathway (8). The enzyme shows a high selectivity for the acetaldehyde acceptor, which would constrain the degradation of ortho-substituted phenylpropionic acids via this pathway. The relaxed specificity for the α-keto acid carbonyl donor offers the possibility of using HKP aldolase for stereospecific carbon-carbon bond formation reactions with nonnatural substrates.  相似文献   

2.
In order to elucidate the structure-activity relationship of griseofulvin (1), (±)-6′-demethyl analog (3), 2′-demethoxy-6′-demethyldihydro analog (4), (±)-dechloro-6′-ethyl analog (5), (±)-dechloro-6′-epi-ethyl analog (6), (±)-6′-ethyl analog (7) and (±)-6′-epi-ethyl analog (8) were synthesized by a Diels-Alder cycloaddition of alkylidene ketones (16, 17, 18, 19 and 20) with modified 1,3-butadienes (21 or 22). Their biological activities were examined against fungi.  相似文献   

3.
The extracellular dispersant of Acinetobacter calcoaceticus A2, referred to as biodispersan, was concentrated by ammonium sulfate precipitation and deproteinized by hot phenol treatment. The active component was an anionic polysaccharide (PS-A2). The specific activity of PS-A2 was approximately three times greater than that of crude biodispersan. PS-A2 had a sedimentation constant of 1.39 S, a diffusion coefficient of 18.8 × 10−8 cm2 s−1, and a partial molar volume of 0.65 cm3 g−1, yielding an average molecular weight of 51,400. Titration of the polymer gave two inflection points: pK1 = 3.1 (1.15 μEq/mg) and pK2 = 8.0 (0.4 μEq/mg). PS-A2 slowly consumed 1.10 μmol of periodate per mg. The 13C nuclear magnetic resonance spectrum of PS-A2 indicated four methyl groups, four carbonyl C atoms, and four signals in the anomeric region (95 to 110 ppm), indicative of the presence of four different monosaccharides. Strong acid hydrolysis of PS-A2 yielded four reducing sugars: glucosamine, a 6-methyl aminohexose, galactosamine uronic acid, and an unidentified amino sugar. Ruthenium red binding to PS-A2 was stoichiometric: 1 molecule of dye bound per 2.0 carboxyl groups.  相似文献   

4.
A DNA fragment that carried the gene (proA) encoding 4-hydroxy-4-methyl-2-oxoglutarate aldolase was cloned from the chromosomal DNA of Pseudomonas ochraceae NGJ1, and the coding region was assigned to the nucleotide sequence based on the N-terminal amino acid sequence of the enzyme purified from the organism. The proA gene was 684 bp long, corresponding to a protein of 227 amino acid residues with a calculated molecular mass of 24,067 Da. The genes encoding a putative transporter and a 4-oxalomesaconate hydratase were upstream, and a 3'-truncated gene encoding 2-pyrone-4,6-dicarboxylate lactonase was downstream from the proA gene in the same orientation on the DNA fragment. The proA gene product was overproduced in Escherichia coli and briefly purified to homogeneity from the crude extract by a two-step purification. The molecular and catalytic properties of the gene product were similar to those of the P. ochraceae enzyme.  相似文献   

5.
The aldolase of Francisella tularensis resembles Class II aldolases in its requirement for divalent ions and its inactivation by metal chelating agents. Cysteine and other reducing agents stimulated the activity of the enzyme.  相似文献   

6.
4-Hydroxy-4-methyl-2-oxoglutarate/4-carboxy-4-hydroxy-2-oxoadipate (HMG/CHA) aldolase from Pseudomonas putida F1 catalyzes the last step of the bacterial protocatechuate 4,5-cleavage pathway. The preferred substrates of the enzyme are 2-keto-4-hydroxy acids with a 4-carboxylate substitution. The enzyme also exhibits oxaloacetate decarboxylation and pyruvate α-proton exchange activity. Sodium oxalate is a competitive inhibitor of the aldolase reaction. The pH dependence of kcat/Km and kcat for the enzyme is consistent with a single deprotonation with pKa values of 8.0 ± 0.1 and 7.0 ± 0.1 for free enzyme and enzyme substrate complex, respectively. The 1.8 Å x-ray structure shows a four-layered α-β-β-α sandwich structure with the active site at the interface of two adjacent subunits of a hexamer; this fold resembles the RNase E inhibitor, RraA, but is novel for an aldolase. The catalytic site contains a magnesium ion ligated by Asp-124 as well as three water molecules bound by Asp-102 and Glu-199′. A pyruvate molecule binds the magnesium ion through both carboxylate and keto oxygen atoms, completing the octahedral geometry. The carbonyl oxygen also forms hydrogen bonds with the guanadinium group of Arg-123, which site-directed mutagenesis confirms is essential for catalysis. A mechanism for HMG/CHA aldolase is proposed on the basis of the structure, kinetics, and previously established features of other aldolase mechanisms.  相似文献   

7.
UDP-galactose 4-epimerase (EC 5.1.3.2) was purified to a homogeneous state from Bifidobacterium bifidutn grown on a glucose medium. The molecular weight was estimated to be about 90,000. The purified enzyme was very stable and 60 % of its initial activity survived three months of storage at 4°C even at a low protein concentration (0.2 mg/ml). The optimum pH was 9.0, and the Km values for UDP-galactose and UDP-glucose were 5.4 × 10-4 M and 1.4×10 -3 M. UDP was a competitive inhibitor. The enzyme activity was stimulated by various sugar phosphates, but was slightly inhibited by fructose 1,6-diphosphate (FDP). A high concentration of galactose or glucose, which had no effect by itself, inhibited the activity in combination with UMP. The inhibition by FDP was also enhanced by combination with UMP.  相似文献   

8.
9.
Although the existence of blaOXA-23 is reported in various parts of the world, the product of blaOXA-23 gene, OXA-23, has not been purified and its kinetic properties are not known. In this study, OXA-23 of Acinetobacter baumannii isolated from Kocaeli University intensive care unit was characterized after purification using recombinant methods. Preliminary results showed that conventional protein purification methods were not effective for purification of OXA-23. Therefore, OXA-23 was fused to maltose-binding protein of Escherichia coli, the fused protein was expressed and purified to homogeneity. Kinetic properties of the pure protein were then studied with substrates e.g., imipenem, meropenem, cefepime, ceftazidime, ampicilline, piperacillin, penicillin G, and nitrocefin. Also clavulanic acid, tazobactam, and sulbactam concentrations that inhibit 50% of OXA-23 enzyme activity were calculated. Modelling of OXA-23 revealed its ionic surface structure, conformation in the fused form and its topology allowing us to make predictions for OXA-23 substrate specificity.  相似文献   

10.
Abstract Membrane-bound l -lactate dehydrogenase has been purified almost to homogeneity from Acinetobacter calcoaceticus . The enzyme is an oligomeric protein of sub-unit M r 40 000 containing non-covalently bound FMN as a prosthetic group. Purified l -lactate dehydrogenase has an apparent K m of 83 μM for l -lactate but has no activity with, and is not inhibited by, d -lactate. The enzyme is strongly inhibited by HgCl2, but other thiol reagents and metal-chelating compounds have little or no effect upon its activity.  相似文献   

11.
Cells of Acinetobacter calcoaceticus contain a constitutive periplasmic metalloproteinase showing similar properties as the periplasmic metalloproteinase of Escherichia coli. The periplasmic proteinase of A. calcoaceticus was purified, starting from periplasm, by ammonium sulfate precipitation, hydrophobic interaction chromatography and chromatofocusing up to the homogeneity of the enzyme in SDS-electrophoresis with a yield of 6.7% and a purification factor of 417. The enzyme has a molecular mass of 108000 (gel filtration) or 112000 (native electrophoresis), and consists of four identical subunits with a molecular mass of 27 000 (SDS-electrophoresis). The purified enzyme degrades preferentially polypeptides such as glucagon and insulin. Larger proteins are accepted as substrates to a considerably lower extent. All tested synthetic substrates with trypsin, chymotrypsin, elastase and thermolysin specificity were not cleaved. Therefore, the described enzyme was designated “insulin-cleaving proteinase” (ICP).  相似文献   

12.
The effect of insulin on polypeptide chain elongation was examined in soleus muscles isolated from 18 hour-fasted mice. Treatment with insulin for 1 hour increased the elongation rate, which was estimated by the half-transit time. This suggests that insulin stimulated protein synthesis by modifying the elongation rate in addition to the initiation rate.  相似文献   

13.
Benzaldehyde dehydrogenase I was purified from Acinetobacter calcoaceticus by DEAE-Sephacel, phenyl-Sepharose and f.p.l.c. gel-filtration chromatography. The enzyme was homogeneous and completely free from the isofunctional enzyme benzaldehyde dehydrogenase II, as judged by denaturing and non-denaturing polyacrylamide-gel electrophoresis. The subunit Mr value was 56,000 (determined by SDS/polyacrylamide-gel electrophoresis). Estimations of the native Mr value by gel-filtration chromatography gave values of 141,000 with a f.p.l.c. Superose 6 column, but 219,000 with Sephacryl S300. Chemical cross-linking of the enzyme subunits indicated that the enzyme is tetrameric. Benzaldehyde dehydrogenase I was activated more than 100-fold by K+, Rb+ and NH4+, and the apparent Km for K+ was 11.2 mM. The pH optimum in the presence of K+ was 9.5 and the pI of the enzyme was 5.55. The apparent Km values for benzaldehyde and NAD+ were 0.69 microM and 96 microM respectively, and the maximum velocity was approx. 110 mumol/min per mg of protein. Various substituted benzaldehydes were oxidized at significant rates, and NADP+ was also used as cofactor, although much less effectively than NAD+. Benzaldehyde dehydrogenase I had an NAD+-activated esterase activity with 4-nitrophenol acetate as substrate, and the dehydrogenase activity was inhibited by a range of thiol-blocking reagents. The absorption spectrum indicated that there was no bound cofactor or prosthetic group. Some of the properties of the enzyme are compared with those of other aldehyde dehydrogenases, specifically the very similar isofunctional enzyme benzaldehyde dehydrogenase II from the same organism.  相似文献   

14.
A lipase-producing bacterium, Acinetobacter calcoacetius LP009, was isolated from raw milk. The optimum conditions for growth and lipase production by A. calcoaceticus LP009 were 15 degrees C with shaking at 200 rpm in LB supplemented with 1.0% (v/v) Tween 80. The crude lipase was purified to homogeneous state by ultrafiltration and gel filtration chromatography on Sephadex G-100. Its molecular weight determined by SDS-PAGE was 23 kDa and it exhibited maximum activity at pH 7.0 and 50 degrees C. It was stable over the pH range of 4.0 to 8.0 and at temperatures lower than 45 degrees C. It was a metalloenzyme that is positionally non-specific and had the ability to improve fat hydrolysis in soybean meal and in premixed animals feed.  相似文献   

15.
鲍曼不动杆菌烈性噬菌体的分离与纯化   总被引:1,自引:0,他引:1  
梁莉  杨洪江  金鑫 《生物学杂志》2010,27(4):88-90,93
利用柱层析方法,纯化鲍曼不动杆菌(Acinetobacter baumannii)烈性噬菌体AB1。首先采用聚乙二醇6000沉淀方法,初步分离裂解液中的噬菌体,噬菌体纯度由6.1×1010 pfu/mg提高到37×1010 pfu/mg,噬菌体回收率为58.8%,蛋白质去除率为90.6%;噬菌体粗提样品经Sepharose 4B凝胶过滤层析柱进一步纯化,纯度提高到73×1010 pfu/mg,噬菌体回收率为95.7%,蛋白质去除率为48.1%;收集的噬菌体样品最后经DEAE-52阴离子交换层析柱处理,噬菌体纯度为40×1010 pfu/mg,回收率为50.8%,蛋白去除率15.6%。内毒素分析结果显示,Sepharose 4B凝胶过滤层析纯化的噬菌体样品中,内毒素含量为443.8 EU/mg,而DEAE-52阴离子交换层析纯化的噬菌体样品中,内毒素含量为544.4 EU/mg。实验结果显示,PEG沉淀方法与Sepharose 4B凝胶过滤方法能够有效地提高噬菌体纯度,而DEAE-52阴离子交换层析则不能提高噬菌体的纯度,也无法有效地去除样品中的内毒素。  相似文献   

16.
Previous studies of potato varieties indicated that changes during cooking could be mathematically described and that some chemical components and the cell size may influence the cooking behavior. To find out whether the same principles can be adopted for other root vegetables, the cooking behavior of three other low-starch root vegetables were investigated and the results compared. Slices (6 mm thick and 30 mm diameter) were treated in water at 100°C. Mathematical expressions were assessed, and coefficients were determined to describe the kinetic behavior of the products. The cell size and pectin content of the raw materials determined the cooking characteristics. Texture development could be predicted by shear force measurements.  相似文献   

17.
2-Ketogluconate reductase (2KGR) from the cell free extract of Gluconobacter liquefaciens (IFO 12388) was purified about 1000-fold by a procedure involving ammonium sulfate fractionation and column chromatographies using DEAE-cellulose, hydroxylapatite, and Sephadex gel The purified enzyme gave a single band on polyacrymamide gel electrophoresis. NADP was specifically required for the oxidation reaction of gluconic acid. Using gel filtration a molecular weight of about 110,000 was estimated for the enzyme. The pH optimum for the oxidation of gluconic acid (GA) to 2-ketogluconic acid (2KGA) by the enzyme was 10.5 and for the reduction of 2KGA was 6.5. The optimum temperature of the enzyme was 50 C for both reactions of oxidation and reduction. The enzyme was stable at pH between 5.0 and 11.0 and at temperature under 50°C, The enzyme activity was strongly inhibited with p-chloromercuribenzoate and mercury ions, but remarkably stimulated by manganese ions (1×10?3 m). Km value of the enzyme for GA was 1.3×10?2 m and for 2KGA was 6.6×10?3 m. Km values for NADP and NADPH2 were 1.25×10?5 and 1.52×10?5 m respectively.  相似文献   

18.
Acinetobacter calcoaceticus ATCC 23055 produces a large amount of 1,3-diaminopropane under normal growth conditions. The enzyme responsible, L-2,4-diaminobutyrate (DABA) decarboxylase (EC 4.1.1.-), was purified to electrophoretic homogeneity from this bacterium. The native enzyme had an M(r) of approximately 108,000, with a pI of 5.0, and was a dimer composed of identical or nearly identical subunits with apparent M(r) 53,000. The enzyme showed hyperbolic kinetics with a Km of 1.59 mM for DABA and 14.6 microM for pyridoxal 5'-phosphate as a coenzyme. The pH optimum was in the range 8.5-8.75, and Ca2+ gave a much higher enzyme activity than Mg2+ as a cationic cofactor. N-gamma-Acetyl-DABA, 2,3-diaminopropionic acid, ornithine and lysine were inert as substrates. The enzyme was different in subunit structure, N-terminal amino acid sequence and immunoreactivity from the DABA decarboxylase of Vibrio alginolyticus previously described.  相似文献   

19.
王志清  庄茂辛 《蛇志》1999,11(4):12-17
目的 从广西产中华眼镜蛇毒 ( N aja naja atra)中分离了一种新的磷脂酶 A2 ( PLA2 ) ,并研究其性质。 方法 磷脂酶 A2 ( PLA2 )的分离纯化采用 CM5 2、CM-Sepharose CL-6 B离子交换柱和 SepharoseCL-6 B凝胶柱 ,磷脂酶 A2 ( PLA2 )的性质采用经典方法进行。 结果 分离后的磷脂酶 A2 经过聚丙烯酰胺凝胶电泳、SDS-聚丙烯酰胺凝胶电泳和 HPLC检测 ,其为单一组分。SDS-PAGE测定它的分子量为1 5 0 0 0± 1 0 0 0。等电聚焦测得它的等电点为 6 .3。它的最适温度为 5 5℃ ,最适 p H值为 8.5。氨基酸成分分析表明该酶由 1 2 6个氨基酸组成 ,以 Asp、Ala、Gly、Cys居多。Fe3 、Zn2 、EDTA对其酶活力起抑制作用 ;K 、Ca2 、去垢剂对其活力起促进作用。荧光光谱分析表明该酶的色氨酸残基、组氨酸残基可能位于分子表面。药理实验表明 :该酶具有抗胰蛋白酶作用、抗凝血作用、间接溶血作用以及对青蛙具有心脏毒性。 结论 广西产中华眼镜蛇毒磷脂酶 A2 与其它来源的 PLA2 同源性高 ,但性质不尽相同。  相似文献   

20.
担子菌漆酶的分离纯化及其性质研究   总被引:26,自引:2,他引:26  
采用硫酸铵盐析、DEAE纤维素柱层析、Phenyl SepharoseTM6 Fast Flow疏水柱层析等方法,得到电泳纯的漆酶同工酶Lac1,纯化倍数为318.4,活力回收率为18.6%。用SDSPAGE测得该酶分子量为60.3kD,而经质谱分析为55.94kD。最适反应温度为65℃,最适反应pH值为2.2~2.8,酶的等电点pI(室温)为4.02,其N末端序列为AIGPVTDL,用硫酸酚法测得其含糖量为49.2%。25℃条件下,以ABTS(2,2'azinobis(3ethylbenzthiazoline6sulphonate)为底物的Km为17.5μmol/L。该酶在45℃,pH3.0~9.5较稳定。Cu2+对酶活有明显的促进作用,Fe2+完全抑制酶的活性,Mn2+和Ag对酶活无明显影响。DTT(Dithiothreitol,二硫苏糖醇)和NaN3完全抑制酶的活性。Koshland试剂对漆酶的活力影响比较大,色氨酸可能是酶活力的必需基团。  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