首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Alpha-ketoglutarate-dependent (R)-dichlorprop dioxygenase (RdpA) and alpha-ketoglutarate-dependent (S)-dichlorprop dioxygenase (SdpA), which are involved in the degradation of phenoxyalkanoic acid herbicides in Sphingomonas herbicidovorans MH, were expressed and purified as His6-tagged fusion proteins from Escherichia coli BL21(DE3)(pLysS). RdpA and SdpA belong to subgroup II of the alpha-ketoglutarate-dependent dioxygenases and share the specific motif HXDX(24)TX(131)HX(10)R. Amino acids His-111, Asp-113, and His-270 and amino acids His-102, Asp-104, and His 257 comprise the 2-His-1-carboxylate facial triads and were predicted to be involved in iron binding in RdpA and SdpA, respectively. RdpA exclusively transformed the (R) enantiomers of mecoprop [2-(4-chloro-2-methylphenoxy)propanoic acid] and dichlorprop [2-(2,4-dichlorophenoxy)propanoic acid], whereas SdpA was specific for the (S) enantiomers. The apparent Km values were 99 microM for (R)-mecoprop, 164 microM for (R)-dichlorprop, and 3 microM for alpha-ketoglutarate for RdpA and 132 microM for (S)-mecoprop, 495 microM for (S)-dichlorprop, and 20 microM for alpha-ketoglutarate for SdpA. Both enzymes had high apparent Km values for oxygen; these values were 159 microM for SdpA and >230 microM for RdpA, whose activity was linearly dependent on oxygen at the concentration range measured. Both enzymes had narrow cosubstrate specificity; only 2-oxoadipate was able to replace alpha-ketoglutarate, and the rates were substantially diminished. Ferrous iron was necessary for activity of the enzymes, and other divalent cations could not replace it. Although the results of growth experiments suggest that strain MH harbors a specific 2,4-dichlorophenoxyacetic acid-converting enzyme, tfdA-, tfdAalpha-, or cadAB-like genes were not discovered in a screening analysis in which heterologous hybridization and PCR were used.  相似文献   

2.
α-Amylase produced by Bacillus licheniformis CUMC305 was purified 212-fold with a 42% yield through a series of four steps. The purified enzyme was homogeneous as shown by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and discontinuous gel electrophoresis. The purified enzyme showed maximal activity at 90°C and pH 9.0, and 91% of this activity remained at 100°C. The enzyme retained 91, 79, and 71% maximal activity after 3 h of treatment at 60°C, 3 h at 70°C, and 90 min at 80°C, respectively, in the absence of substrate. On the contrary, in the presence of substrate (soluble starch), the α-amylase enzyme was fully stable after a 4-h incubation at 100°C. The enzyme showed 100% stability in the pH range 7 to 9; 95% stability at pH 10; and 84, 74, 68, and 50% stability at pH values of 6, 5, 4, and 3, respectively, after 18 h of treatment. The activation energy for this enzyme was calculated as 5.1 × 105 J/mol. The molecular weight was estimated to be 28,000 by sodium dodecyl sulfate-gel electrophoresis. The relative rates of hydrolysis of soluble starch, amylose, amylopectin, and glycogen were 1.27, 1.8, 1.94, and 2.28 mg/ml, respectively. Vmax values for hydrolysis of these substrates were calculated as 0.738, 1.08, 0.8, and 0.5 mg of maltose/ml per min, respectively. Of the cations, Na+, Ca2+, and Mg2+, showed stimulatory effect, whereas Hg2+, Cu2+, Ni2+, Zn2+, Ag+, Fe2+, Co2+, Cd2+, Al3+, and Mn2+ were inhibitory. Of the anions, azide, F, SO32−, SO43−, S2O32−, MoO42−, and Wo42− showed an excitant effect. p-Chloromercuribenzoic acid and sodium iodoacetate were inhibitory, whereas cysteine, reduced glutathione, thiourea, β-mercaptoethanol, and sodium glycerophosphate afforded protection to enzyme activity. α-Amylase was fairly resistant to EDTA treatment at 30°C, but heating at 90°C in presence of EDTA resulted in the complete loss of enzyme activity, which could be recovered partially by the addition of Cu2+ and Fe2+ but not by the addition of Ca2+ or any other divalent ions.  相似文献   

3.
Two α-ketoglutarate-dependent dioxygenases carrying enantiospecific activity for the etherolytic cleavage of racemic phenoxypropionate herbicides [(RS)-2-(2,4-dichlorophenoxy)propionate and (RS)-2-(4-chloro-2-methylphenoxy)propionate] from Delftia acidovorans MC1 were characterized with respect to protein and sequence data. The (S)-phenoxypropionate/α-ketoglutarate-dioxygenase (SdpA) appeared as a monomeric enzyme with a molecular weight of 32 kDa in the presence of SDS. N-terminal sequences revealed relationship to α-ketoglutarate-dependent taurine dioxygenase (TauD) and to 2,4-dichlorophenoxyacetate/α-ketoglutarate-dioxygenase (TfdA). The (R)-phenoxypropionate/α-ketoglutarate-dioxygenase (RdpA) referred to 36 kDa in the presence of SDS and to 108 kDa under native conditions. Internal sequences of fragments obtained after digestion made evident relationship to TfdA and TauD. Two-dimensional electrophoretic separation resulted in the resolution of up to 3 individual spots with almost identical molecular weights but different isoelectric points with both RdpA and SdpA. The structural differences of these isoenzyme forms are not yet clear.  相似文献   

4.
The etherolytic cleavage of phenoxyalkanoic acids in various bacteria is catalyzed by an α-ketoglutarate-dependent dioxygenase. In this reaction, the electron acceptor is oxidatively decarboxylated to succinate, whereas the proper substrate is cleaved by forming the oxidized alkanoic acid and the phenolic intermediate. The necessity of regenerating α-ketoglutarate and the consequences for the overall metabolism were investigated in a theoretical study. It was found that the dioxygenase mechanism is accompanied by a significant loss of carbon amounting to up to 62.5% in the assimilatory branch, thus defining the upper limit of carbon conversion efficiency. This loss in carbon is almost compensated for in comparison to a monooxygenase-catalyzed initial step when the dissimilatory efforts of the entire metabolism are included: the yield coefficients become similar. The α-ketoglutarate-dependent dioxygenase mechanism has more drastic consequences for microorganisms which are restricted in their metabolism to the first step of phenoxyalkanoate degradation by excreting the phenolic intermediate as a dead-end product. In the case of phenoxyacetate derivatives, the cleavage reaction would quickly cease due to the exhaustion of α-ketoglutarate and no growth would be possible. With the cleavage products of phenoxypropionate and phenoxybutyrate herbicides, i.e., pyruvate and succinate(semialdehyde), respectively, as the possible products, the regeneration of α-ketoglutarate will be guaranteed for stoichiometric reasons. However, the maintenance of the cleavage reaction ought to be restricted due to physiological factors owing to the involvement of other metabolic reactions in the pool of metabolites. These effects are discussed in terms of a putative recalcitrance of these compounds.  相似文献   

5.
Constitutive beta-glucosidases from Saccharomyces fragilis (Y-18) and S. dobzhanskii (Y-19) precipitated at the same concentration of ammonium sulfate. The partially purified enzymes had similar activation energies, molecular weights, affinities for certain natural and synthetic beta-glucosides, and optimal pH values for substrate hydrolysis, and they were stable over approximately the same pH range. The enzymes, however, could be clearly distinguished by other criteria. Affinities of the synthetic, sulfur-containing beta-glucosides for Y-18 enzyme were many times greater than for Y-19 enzyme. The latter enzyme was more resistant to heat. The two enzymes eluted from diethylaminoethyl cellulose at different concentrations of sodium chloride. In precipitin tests, homologous enzyme-antisera systems were highly specific. The beta-glucosidase synthesized by a hybrid, S. fragilis x S. dobzhanskii (Y-42), was unique. Characterization of this enzyme produced values which were intermediate to those for the enzymes from the parental yeast strains. Heat-inactivation slopes and Lineweaver-Burk plots for the Y-42 enzyme were anomalous. It is suggested that hydrolytic activity in Y-42 preparations is due to a spectrum of hybrid enzyme molecules composed of varying amounts of two distinct polypeptides. It is further suggested that these polypeptides may be identical to those synthesized by the parental Y-18 and Y-19 yeast strains.  相似文献   

6.
α-Ionone, α-methylionone, and α-isomethylionone were converted by Aspergillus niger JTS 191. The individual bioconversion products from α-ionone were isolated and identified by spectrometry and organic synthesis. The major products were cis-3-hydroxy-α-ionone, trans-3-hydroxy-α-ionone, and 3-oxo-α-ionone. 2,3-Dehydro-α-ionone, 3,4-dehydro-β-ionone, and 1-(6,6-dimethyl-2-methylene-3-cyclohexenyl)-buten-3-one were also identified. Analogous bioconversion products from α-methylionone and α-isomethylionone were also identified. From results of gas-liquid chromatographic analysis during the fermentation, we propose a metabolic pathway for α-ionones and elucidation of stereochemical features of the bioconversion.  相似文献   

7.
Two α-glucosidase-encoding genes (agl1 and agl2) from Bifidobacterium breve UCC2003 were identified and characterized. Based on their similarity to characterized carbohydrate hydrolases, the Agl1 and Agl2 enzymes are both assigned to a subgroup of the glycosyl hydrolase family 13, the α-1,6-glucosidases (EC 3.2.1.10). Recombinant Agl1 and Agl2 into which a His12 sequence was incorporated (Agl1His and Agl2His, respectively) exhibited hydrolytic activity towards panose, isomaltose, isomaltotriose, and four sucrose isomers—palatinose, trehalulose, turanose, and maltulose—while also degrading trehalose and, to a lesser extent, nigerose. The preferred substrates for both enzymes were panose, isomaltose, and trehalulose. Furthermore, the pH and temperature optima for both enzymes were determined, showing that Agl1His exhibits higher thermo and pH optima than Agl2His. The two purified α-1,6-glucosidases were also shown to have transglycosylation activity, synthesizing oligosaccharides from palatinose, trehalulose, trehalose, panose, and isomaltotriose.  相似文献   

8.
Sphingomonas herbicidovorans MH was able to completely degrade both enantiomers of the chiral herbicide dichlorprop [(RS)-2-(2,4-dichlorophenoxy)propanoic acid], with preferential degradation of the (S) enantiomer over the (R) enantiomer. These results are in agreement with the recently reported enantioselective degradation of mecoprop [(RS)-2-(4-chloro-2-methylphenoxy)propanoic acid] by this bacterium (C. Zipper, K. Nickel, W. Angst, and H.-P. E. Kohler, Appl. Environ. Microbiol. 62:4318–4322, 1996). Uptake of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D (2,4-dichlorophenoxyacetic acid) was inducible. Initial uptake rates of cells grown on the respective substrate showed substrate saturation kinetics with apparent affinity constants (Kt) of 108, 93, and 117 μM and maximal velocities (Vmax) of 19, 10, and 21 nmol min−1 mg of protein−1 for (R)-dichlorprop, (S)-dichlorprop, and 2,4-D, respectively. Transport of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D was completely inhibited by various uncouplers and by nigericin but was only marginally inhibited by valinomycin and by the ATPase inhibitor N,N′-dicyclohexylcarbodiimine. Experiments on the substrate specificity of the putative transport systems revealed that (R)-dichlorprop uptake was inhibited by (R)-mecoprop but not by (S)-mecoprop, (S)-dichlorprop, or 2,4-D. On the other hand, the (S)-dichlorprop transport was inhibited by (S)-mecoprop but not by (R)-mecoprop, (R)-dichlorprop, or 2,4-D. These results provide evidence that the first step in the degradation of dichlorprop, mecoprop, and 2,4-D by S. herbicidovorans is active transport and that three inducible, proton gradient-driven uptake systems exist: one for (R)-dichlorprop and (R)-mecoprop, another for (S)-dichlorprop and (S)-mecoprop, and a third for 2,4-D.  相似文献   

9.
1. The presence of beta-galactosidase (EC 3.2.1.23) in an acetic acid extract of ram testis is reported. Some properties of the crude enzyme preparation were studied. 2. The purification of beta-acetylglucosaminase (EC 3.2.1.30) and of beta-galactosidase from the ram-testis extract by ammonium sulphate precipitation and chromatography on a CM-cellulose column is described. 3. The final purifications of the separated enzymes achieved were for the beta-acetylglucosaminase 35 times and for the beta-galactosidase 99 times. 4. The possibility of using DEAE-cellulose and Sephadex G-200 to purify the enzymes was investigated.  相似文献   

10.
采用离子交换层析和凝胶过滤层析对鳞杯伞子实体中的α-半乳糖苷酶进行纯化,得到了一种分子量为50 kDa的α-半乳糖苷酶,命名为CSG。纯化后的CSG纯化倍数为891.46倍,比活力为54.78 U/mg,得率为0.71%。通过BLAST比对液相色谱-串联质谱(LC-MS/MS)获得其肽段,发现其为GH27家族的α-半乳糖苷酶。CSG的最适pH为3.0,最适温度为50 ℃。在酸性范围pH 2.2-7.0和温度范围4-30 ℃有较好的稳定性。Mn2+、Cd2+、Cu2+对CSG有较强的抑制作用。半乳糖和蜜二糖对CSG的抑制类型为混合型抑制。化学修饰剂N-溴代琥珀酰亚胺显著降低CSG的活力,碳二亚胺对CSG具有显著的激活作用。该酶具有良好的蛋白酶抗性,且对棉子糖家族寡糖(RFOs)、瓜尔豆胶和赤槐豆胶均表现出良好的水解作用。  相似文献   

11.
Sphingomonas paucimobilis B90A contains two variants, LinA1 and LinA2, of a dehydrochlorinase that catalyzes the first and second steps in the metabolism of hexachlorocyclohexanes (R. Kumari, S. Subudhi, M. Suar, G. Dhingra, V. Raina, C. Dogra, S. Lal, J. R. van der Meer, C. Holliger, and R. Lal, Appl. Environ. Microbiol. 68:6021-6028, 2002). On the amino acid level, LinA1 and LinA2 were 88% identical to each other, and LinA2 was 100% identical to LinA of S. paucimobilis UT26. Incubation of chiral α-hexachlorocyclohexane (α-HCH) with Escherichia coli BL21 expressing functional LinA1 and LinA2 S-glutathione transferase fusion proteins showed that LinA1 preferentially converted the (+) enantiomer, whereas LinA2 preferred the (−) enantiomer. Concurrent formation and subsequent dissipation of β-pentachlorocyclohexene enantiomers was also observed in these experiments, indicating that there was enantioselective formation and/or dissipation of these enantiomers. LinA1 preferentially formed (3S,4S,5R,6R)-1,3,4,5,6-pentachlorocyclohexene, and LinA2 preferentially formed (3R,4R,5S,6S)-1,3,4,5,6-pentachlorocyclohexene. Because enantioselectivity was not observed in incubations with whole cells of S. paucimobilis B90A, we concluded that LinA1 and LinA2 are equally active in this organism. The enantioselective transformation of chiral α-HCH by LinA1 and LinA2 provides the first evidence of the molecular basis for the changed enantiomer composition of α-HCH in many natural environments. Enantioselective degradation may be one of the key processes determining enantiomer composition, especially when strains that contain only one of the linA genes, such as S. paucimobilis UT26, prevail.  相似文献   

12.
l-Methionine γ-lyase (EC 4.4.1.11) was purified to homogeneity from Brevibacterium linens BL2, a coryneform bacterium which has been used successfully as an adjunct bacterium to improve the flavor of Cheddar cheese. The enzyme catalyzes the α,γ elimination of methionine to produce methanethiol, α-ketobutyrate, and ammonia. It is a pyridoxal phosphate-dependent enzyme, with a native molecular mass of approximately 170 kDa, consisting of four identical subunits of 43 kDa each. The purified enzyme had optimum activity at pH 7.5 and was stable at pHs ranging from 6.0 to 8.0 for 24 h. The pure enzyme had its highest activity at 25°C but was active between 5 and 50°C. Activity was inhibited by carbonyl reagents, completely inactivated by dl-propargylglycine, and unaffected by metal-chelating agents. The pure enzyme had catalytic properties similar to those of l-methionine γ-lyase from Pseudomonas putida. Its Km for the catalysis of methionine was 6.12 mM, and its maximum rate of catalysis was 7.0 μmol min−1 mg−1. The enzyme was active under salt and pH conditions found in ripening Cheddar cheese but susceptible to degradation by intracellular proteases.

Methanethiol is associated with desirable Cheddar-type sulfur notes in good-quality Cheddar cheese (2, 27). The mechanism for the production of methanethiol in cheese is unknown, but it is linked to the catabolism of methionine (1, 15). l-Methionine γ-lyase (EC 4.4.1.11; MGL), also known as methionase, l-methionine γ-demethiolase, and l-methionine methanethiollyase (deaminating), is a pyridoxal phosphate (PLP)-dependent enzyme that catalyzes the direct conversion of l-methionine to α-ketobutyrate, methanethiol, and ammonia by an α,γ-elimination reaction (26). It does not catalyze the conversion of d enantiomers (2426). MGL in Pseudomonas putida is a multifunctional enzyme system since it catalyzes the α,γ- and α,β-elimination reactions of methionine and its derivatives (24). In addition, the enzyme also catalyzes the β-replacement reactions of sulfur amino acids (24). Since its discovery in Escherichia coli and Proteus vulgaris by Onitake (19), this enzyme has been found in various bacteria and is regarded as a key enzyme in the bacterial metabolism of methionine. However, this enzyme has not been purified to homogeneity from any food-grade microorganisms.MGL is widely distributed in bacteria, especially in pseudomonads, and is induced by the addition of l-methionine to the culture medium (9, 28). The enzyme has been purified from Pseudomonas putida (25), Aeromonas sp. (26), Clostridium sporogenes (11), and Trichomonas vaginalis (16) and partially purified from and characterized for Brevibacterium linens NCDO 739 (4).B. linens is a nonmotile, non-spore-forming, non-acid-fast, gram-positive coryneform bacterium normally found on the surfaces of Limburger and other Trappist-type cheeses. This organism tolerates salt concentrations ranging between 8 and 20% and is capable of growing in a broad pH range from 5.5 to 9.5, with an optimum pH of 7.0 (20). In Trappist-type cheeses, brevibacteria depend on Saccharomyces cerevisiae to metabolize lactate, which increases the pH of the curd, as well as to produce growth factors that are important for their growth (20). Interest in B. linens has focused around its ability to produce an extracellular protease, which has recently been isolated (21), and its ability to produce high levels of methanethiol (3, 9, 10, 22).B. linens produces various sulfur compounds, including methanethiol, that are thought to be important in Cheddar-like flavor and aroma (3, 9, 10, 22). Ferchichi et al. (9) suggested that MGL is responsible for the methanethiol-producing capability of B. linens but did not provide definitive evidence. Weimer et al. (28) proposed that B. linens BL2 is responsible for Cheddar-type flavor development in low-fat cheese, but again conclusive evidence was lacking. In this study, MGL was purified to homogeneity from B. linens BL2 and its physical and chemical properties were examined.  相似文献   

13.
We report the isolation and physiological characterization of novel, psychrophilic, iron-oxidizing bacteria (FeOB) from low-temperature weathering habitats in the vicinity of the Juan de Fuca deep-sea hydrothermal area. The FeOB were cultured from the surfaces of weathered rock and metalliferous sediments. They are capable of growth on a variety of natural and synthetic solid rock and mineral substrates, such as pyrite (FeS2), basalt glass (~10 wt% FeO), and siderite (FeCO3), as their sole energy source, as well as numerous aqueous Fe substrates. Growth temperature characteristics correspond to the in situ environmental conditions of sample origin; the FeOB grow optimally at 3 to 10°C and at generation times ranging from 57 to 74 h. They are obligate chemolithoautotrophs and grow optimally under microaerobic conditions in the presence of an oxygen gradient or anaerobically in the presence of nitrate. None of the strains are capable of using any organic or alternate inorganic substrates tested. The bacteria are phylogenetically diverse and have no close Fe-oxidizing or autotrophic relatives represented in pure culture. One group of isolates are γ-Proteobacteria most closely related to the heterotrophic bacterium Marinobacter aquaeolei (87 to 94% sequence similarity). A second group of isolates are α-Proteobacteria most closely related to the deep-sea heterotrophic bacterium Hyphomonas jannaschiana (81 to 89% sequence similarity). This study provides further evidence for the evolutionarily widespread capacity for Fe oxidation among bacteria and suggests that FeOB may play an unrecognized geomicrobiological role in rock weathering in the deep sea.  相似文献   

14.
A thermostable amylase, possibly a β-amylase from Thermoactinomyces sp. no. 2 isolated from soil, is reported. The enzyme was purified 36-fold by acetone precipitation, ion-exchange chromatography, and Sephadex G-200 gel filtration, and the molecular weight was estimated at 31,600. The enzyme was characterized by demonstration of optimum activity at 60°C and pH 7 and by retention of 70% activity at 70°C (30 min). It was stimulated by Mn2+ and Fe2+ but strongly inhibited by Hg2+. Maltose was the only detectable product of hydrolysis of starches and was quantitatively highest in plantain starch hydrolysate.  相似文献   

15.
There is increasing evidence that a fine-tuned integrin cross talk can generate a high degree of specificity in cell adhesion, suggesting that spatially and temporally coordinated expression and activation of integrins are more important for regulated cell adhesive functions than the intrinsic specificity of individual receptors. However, little is known concerning the molecular mechanisms of integrin cross talk. With the use of beta(1)-null GD25 cells ectopically expressing the beta(1)A integrin subunit, we provide evidence for the existence of a cross talk between beta(1) and alpha(V) integrins that affects the ratio of alpha(V)beta(3) and alpha(V)beta(5) integrin cell surface levels. In particular, we demonstrate that a down-regulation of alpha(V)beta(3) and an up-regulation of alpha(V)beta(5) occur as a consequence of beta(1)A expression. Moreover, with the use of GD25 cells expressing the integrin isoforms beta(1)B and beta(1)D, as well as two beta(1) cytoplasmic domain deletion mutants lacking either the entire cytoplasmic domain (beta(1)TR) or only its "variable" region (beta(1)COM), we show that the effects of beta(1) over alpha(V) integrins take place irrespective of the type of beta(1) isoform, but require the presence of the "common" region of the beta(1) cytoplasmic domain. In an attempt to establish the regulatory mechanism(s) whereby beta(1) integrins exert their trans-acting functions, we have found that the down-regulation of alpha(V)beta(3) is due to a decreased beta(3) subunit mRNA stability, whereas the up-regulation of alpha(V)beta(5) is mainly due to translational or posttranslational events. These findings provide the first evidence for an integrin cross talk based on the regulation of mRNA stability.  相似文献   

16.
The purification of staphylococcal β-hemolysin was accomplished by the successive use of three protein fractionation methods. The first method employed was a double precipitation with the use of ammonium sulfate at 65% saturation. The second phase of purification used Sephadex G-100 column fractionation. The third phase utilized either carboxymethyl cellulose or diethylaminoethyl cellulose fractionation. The last two fractionation methods both resulted in the separation of a relatively high concentration of cationic hot-cold lysin and a low concentration of anionic hot-cold lysin. Because of the low concentration of the anionic component, its purity could not be assessed. However, the purity of the cationic component was demonstrated by immunodiffusion, microimmunoelectrophoresis, and by disc polyacrylamide gel electrophoresis. In addition, antisera against purified cationic β-hemolysin yielded one line of precipitate when tested against the original crude β-hemolysin. The purified cationic β-hemolysin was stable in the lyophilized state. Crude β-hemolysin was dermonecrotic, whereas purified cationic β-hemolysin was not dermonecrotic even after Mg++ activation.  相似文献   

17.
18.
19.
Twelve species of Streptomyces that formerly belonged to the genus Chainia were screened for the production of xylanase and cellulase. One species, Streptomyces roseiscleroticus (Chainia rosea) NRRL B-11019, produced up to 16.2 IU of xylanase per ml in 48 h. A xylanase from S. roseiscleroticus was purified and characterized. The enzyme was a debranching β-(1-4)-endoxylanase showing high activity on xylan but essentially no activity against acid-swollen (Walseth) cellulose. It had a very low apparent molecular weight of 5,500 by native gel filtration, but its denatured molecular weight was 22,600 by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. It had an isoelectric point of 9.5. The pH and temperature optima for hydrolysis of arabinoxylan were 6.5 to 7.0 and 60°C, respectively, and more than 75% of the optimum enzyme activity was retained at pH 8.0. The xylanase had a Km of 7.9 mg/ml and an apparent Vmax of 305 μmol · min-1 · mg of protein-1. The hydrolysis rate was linear for xylan concentrations of less than 4 mg/ml, but significant inhibition was observed at xylan concentrations of more than 10 mg/ml. The predominant products of arabinoxylan hydrolysis included arabinose, xylobiose, and xylotriose.  相似文献   

20.
A cationic protein of rabbit serum bactericidal for Staphylococcus aureus was purified. The specific activity per unit of protein of the purified staphylocidal preparation was approximately 37,000 times greater than that of the serum from which it was isolated. Similar techniques were used to purify serum beta-lysin active against Bacillus subtilis approximately 24,000 times. The staphylocidal activity cannot be attributed to the same beta-lysin active against B. subtilis, lysozyme, or antibody-complement systems. The concentrations of staphylocidal beta-lysin in the sera of the five mammalian species studied did not correlate with their beta-lysin activities against B. subtilis. The two beta-lysins are similar in that both were heat-stable, sensitive to trypsin digestion, had molecular weights near 6,000, and were found in higher concentrations in serum than in plasma. Furthermore, similar techniques can be used to absorb and elute both substances in highly purified forms using cellulose asbestos filter pads and ion exchange chromatography on carboxymethyl cellulose. In contrast to the beta-lysin against B. subtilis, the staphylocidal beta-lysin was not released from blood platelets, and it was inactive in the presence of heparin, sodium citrate, sodium oxalate, ethylenediaminetetraacetic acid, acidic phospholipids, and acid pH values. A variety of proteins, including those of normal serum, preferentially inhibited the bactericidal activity of staphylocidal beta-lysin but not the beta-lysin against B. subtilis.  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