首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 390 毫秒
1.
The thermal coefficient of expansion of egg lecithin bilayer thickness, αd1, was measured as a function of its cholesterol content up to mole ratio lecithin/cholesterol of 1:1, and over the temperature range 0–40 °C. At all cholesterol contents αd1 changes abruptly at approximately 12 °C indicating a structural transition at this temperature. Above 12 °C, αd1 decreases monotonically from −2·10−3 for pure egg lecithin to −1·10–3 at mole ratio 1:1. Below 12 °C αd1 is walways higher than above 12 °C and shows a sharp, anomalously high value of −6·10−3 at the mole ratio 2:1. The results have been interpreted as the movement of cholesterol into the bilayer or the formation of lecithin-cholesterol “complexes” at temperatures below 12 °C. Similar studies with phosphatidylinositol containing cholesterol showed no structural transition and lysolecithin containing cholesterol behaved differently giving two lamellar phases in equilibrium.  相似文献   

2.
Information is given concerning two standard buffer solutions suitable as pH references in 30, 40, and 50 mass% dimethyl sulfoxide (DMSO)/H2O mixed solvents at subzero temperatures from −20 to 0 °C, with the intention of establishing a pH (designated pH*) scale. The two buffers selected were the ampholytes N,N-bis(2-hydroxyethyl)-2-aminoethane sulfonic acid (“bes”) and N-tris(hydroxymethyl)methylglycine (“tricine”), and the reference standard consisted of equal molal quantities of the buffer and its respective sodium salt. The assignment of pH* values was based on measurements of the emf of cells without liquid junction of the type: Pt;H2(g,1 atm) ¦Bes, Na Besate, NaCl ¦ AgCl;Ag and Pt;H2(g,1 atm) ¦Tricine, Na Tricinate, NaCl ¦AgCl;Ag and the pH* was derived from a determination of K2, the equilibrium constant for the dissociation process (Buffer)±/ai (Buffer) + H+.  相似文献   

3.
The Rhodococcus erythropolis strain (N′4) possesses the ability to convert 4-chloro-3-hydroxybutyronitrile into the corresponding acid. This conversion was determined to be performed by its nitrile hydratase and amidase. Ammonium sulfate fractionation, DEAE ion exchange chromatography, and phenyl chromatography were used to partially purify nitrile hydratase from cell-free extract. A SDS-PAGE showed that the partially purified enzyme had two subunits and gel filtration chromatography showed that it consisted of four subunits of α2β2. The purified enzyme had a high specific activity of 860 U mg−1 toward methacrylonitrile. The enzyme was found to have high activity at low temperature range, with a maximum activity occurring at 25 °C and be stable in the presence of organic acids at higher temperatures. The enzyme exhibited a preference for aliphatic saturated nitrile substrates over aliphatic unsaturated or aromatic ones. It was inhibited by sulfhydryl, oxidizing, and serine protease inhibitors, thus indicating that essential cysteine and serine residues can be found in the active site.The purified nitrile hydratase was able to convert 4-chloro-3-hydroxybutyronitrile into the corresponding amide at 15 °C. GC analysis showed that the initial conversion rate of the reaction was 215 mg substrate consumed min−1 mg−1. This demonstrated that this enzyme could be used in conjunction with a stereoselective amidase to synthesize ethyl (S)-4-chloro-3-hydroxybutyrate, an intermediate for a hypercholesterolemia drug, Atorvastatin.  相似文献   

4.
A novel raw starch degrading α-cyclodextrin glycosyltransferase (CGTase; E.C. 2.4.1.19), produced by Klebsiella pneumoniae AS-22, was purified to homogeneity by ultrafiltration, affinity and gel filtration chromatography. The specific cyclization activity of the pure enzyme preparation was 523 U/mg of protein. No hydrolysis activity was detected when soluble starch was used as the substrate. The molecular weight of the pure protein was estimated to be 75 kDa with SDS-PAGE and gel filtration. The isoelectric point of the pure enzyme was 7.3. The enzyme was most active in the pH range 5.5–9.0 whereas it was most stable in the pH range 6–9. The CGTase was most active in the temperature range 35–50°C. This CGTase is inherently temperature labile and rapidly loses activity above 30°C. However, presence of soluble starch and calcium chloride improved the temperature stability of the enzyme up to 40°C. In presence of 30% (v/v) glycerol, this enzyme was almost 100% stable at 30°C for a month. The Km and kcat values for the pure enzyme were 1.35 mg ml−1 and 249 μM mg−1 min−1, respectively, with soluble starch as the substrate. The enzyme predominantly produced α-cyclodextrin without addition of any complexing agents. The conditions employed for maximum α-cyclodextrin production were 100 g l−1 gelatinized soluble starch or 125 g l−1 raw wheat starch at an enzyme concentration of 10 U g−1 of starch. The α:β:γ-cyclodextrins were produced in the ratios of 81:12:7 and 89:9:2 from gelatinized soluble starch and raw wheat starch, respectively.  相似文献   

5.
(Ph4P)4[Tl4Se16] was prepared hydrothermally in a sealed pyrex tube by the reaction of TlCl, K2Se4 and Ph4PCl in a 1:1:1 molar ratio at 110 °C for one day. The red crystals were obtained in 50% yield. Crystals of (Ph4P)4[Tl4Se16]: triclinic P (No. 2), Z=1, a=12.054(9), b=19.450(10), c=11.799(6) Å, α=104.63(4), β=98.86(6), γ=101.99(6)° and V=2555(3) Å3 at 23 °C, 2θmax=40.0°, μ=120.7 cm−1, Dcalc=2.23. The structure was solved by direct methods. Number of data collected: 5206. Number of unique data having Fo2>3σ(Fo2): 1723. Final R=0.075 and Rw=0.089. [Tl4Se16]4− consists of four, almost already linearly arranged, tetrahedral thallium centers which are coordinated by two chelating Se42−, two bridging Se22− and four bridging Se2− ligands. [Tl4Se16]4− sits on an inversion center and possesses a central {Tl2Se2}2+ planar core. The Tl(1)–Tl(1)′ distance in this core is 3.583(6) Å. These two thallium atoms are then each linked to two cyclic Tl(Se4) fragments via bridging Se22− and Se2− ligands forming Tl2Se(Se2) five-membered rings.  相似文献   

6.
A highly sensitive, kinetically unambiguous assay for α-factor-induced delay of cell passage through the “start” step of cell division in yeast is presented. The assay employs perfusion with periodic microscopy to monitor the bud emergence kinetics on the 20% of cells within an exponentially growing population which exist prior to the α-factor execution point of start. The t1/2 for cell passage through start by this population of cells is 31 min in the absence of α-factor. The inhibition constant, KI, represents the α-factor concentration which produces a 50% inhibition of this rate and is equal to 2×10−10M. A second assay for maximal cell division arrest by α-factor on whole populations of cells is presented. This assay shows a maximum cell division arrest time of 125±5 h at saturating α-factor, and a K50 (that is, an α-factor concentration which produces a half-maximal response) of 2.5×10−8M. Both assays were performed in the effective absence of α-factor inactivation. Values of the dissociation constant KD and total number of receptors per cell which specifically mediate cell division arrest or delay were estimated to be 2.5×10−8M and 104, respectively. These estimates, along with the quantitative dose-response data for division arrest which are presented here, are consistent with each receptor·α-factor complex which is present on the cell at equilibrium producing a 43±10 s delay of cell passage through start. Surprisingly, this number is constant within twofold over the entire range of cellular division arrest responses to α-factor, that is, from a 1.9-fold inhibition of the rate of cell passage through start at 0.17 nM α-factor to a 125±5 h maximum arrest at saturating α-factor concentrations of >170 nM. The possible significance of this observation toward the mechanism of α-factor-induced cell division arrest is discussed.  相似文献   

7.
Reaction of sodium or potassium molybdate and excess malic acid in a wide range of pH values (pH 4.0–7.0) resulted in the isolation of two cis-dioxo-bis(malato)-Mo(VI) complexes, viz. Na3[MoO2H(S-mal)2] and K3[MoO2H(S-mal)2]·H2O (H3mal=malic acid). The sodium complex is also characterized by an X-ray structure analysis, showing that the mononuclear Mo units are linked together via very strong symmetric CO2···H··· O2C-hydrogen bond [2.432(5) Å], forming a polymeric chain. The molybdenum atoms are quasi-octahedrally coordinated by two cis-oxo groups and two bidentate malate ligands via its alkoxy and α-carboxyl groups, while the β-carboxylic and carboxylate groups remain uncomplexed, as the coordination of vicinal carboxylate and alkoxide of homocitrate in FeMo cofactor of nitrogenase. The absolute configuration of the metal center in this S-malato complex is assigned as Λ and the homochirality within the chain is established as a homochiral form ···ΛS–ΛS–ΛS–ΛS···. It is proposed that the chiral configuration of the metal center in wild-type FeMo-co biosynthesis might be induced by the early coordination of the chiral R-homocitric acid, while a mixture of raceme might be obtained in the biosynthesis of NifV FeMo-cofactor. The absolute configuration of wild-type FeMo-cofactor is assigned as ΔR.  相似文献   

8.
An intracellular β-xylosidase from the thermophilic fungus Sporotricum thermophile strain ATCC 34628 was purified to homogeneity by Q-Sepharose and Mono-Q column chromatographies. The protein properties correspond to molecular mass and pI values of 45 kDa and 4.2, respectively. The enzyme is optimally active at pH 7.0 and 50 °C. The purified β-xylosidase is fully stable at pH 6.0–8.0 and temperatures up to 50 °C and retained over 58% of its activity after 1 h at 60 °C. The enzyme hydrolyzes β-1,4-linked xylo-oligosaccharides with chain lengths from 2 to 6, releasing xylose from the non-reducing end, but is inactive against xylan substrates. The apparent Km and Vmax values from p-nitrophenyl β-d-xylopyranoside are 1.1 mM and 114 μmol p-nitrophenol min−1 mg−1, respectively. Alcohols inactivate the enzyme, ethanol at 10% (v/v) yields a 30% decrease of its activity. The enzyme is irreversibly inhibited by 2,3-epoxypropyl β-d-xylobioside while alkyl epoxides derived from d-xylose were not inhibitors of the enzyme. The enzyme catalyses the condensation reaction using high donor concentration, up to 60% (w/v) xylose.  相似文献   

9.
Amniotic fluid from 15–17-day rat fetuses bound estrone and 17β-estradiol specifically. Related steroids such as estriol, 6-ketoestradiol, 17α-estradiol and testosterone were not bound to any significant extent. The apparent Ka for 17β-estradiol was 2.6·108 M at 4°C; 6 nmoles of 17β-estradiol were bound per ml of amniotic fluid. The binding component appears to be α-fetoprotein in that it migrates as an α1-globulin upon polyacrylamide gel electrophoresis and has an isoelectric pH of 4.7 as determined by isoelectric focusing. Furthermore, binding activity was precipitated by antiserum which was shown by immuno-electrophoresis to be specific for α-fetoprotein. Binding activity, partially purified by isoelectric focusing of amniotic fluid, was associated with one of two bands seen by polyacrylamide gel electrophoresis. This band migrated as an α1-globulin.  相似文献   

10.
A novel β-glucosidase from Fusarium proliferatum ECU2042 (FPG) was successfully purified to homogeneity with a 506-fold increase in specific activity. The molecular mass of the native purified enzyme (FPG) was estimated to be approximately 78.7 kDa, with two homogeneous subunits of 39.1 kDa, and the pI of this enzyme was 4.4, as measured by two-dimensional electrophoresis. The optimal activities of FPG occurred at pH 5.0 and 50 °C, respectively. The enzyme was stable at pH 4.0–6.5 and temperatures below 60 °C, and the deactivation energy (Ed) for FPG was 88.6 kJ mo1−1. Moreover, it was interesting to find that although the purified enzyme exhibited a very low activity towards p-nitrophenyl β-d-glucoside (pNPG), and almost no activity towards cellobiose, a relatively high activity was observed on ginsenoside Rg3. The enzyme hydrolyzed the 3-C, β-(1 → 2)-glucoside of ginsenoside Rg3 to produce ginsenoside Rh2, but did not sequentially hydrolyze the β-d-glucosidic bond of Rh2. The Km and Vmax values of FPG for ginsenoside Rg3 were 2.37 mM and 0.568 μmol (h mg protein)−1, respectively. In addition, this enzyme also exhibited significant activities towards various alkyl glucosides, aryl glucosides and several natural glycosides.  相似文献   

11.
The absolute configuration of cis-epoxyjasmone (−)-2, isolated from Trichosporum cutaneum CCT 1903 whole cells, has been unambiguously established as (7S,8R), [α]D20 −29.0° (c 1.3, CHCl3), by a new two step method, using a regioselective epoxide opening as the key step followed by Mosher acid derivatization.  相似文献   

12.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

13.
The NAD-dependent glutamate dehydrogenase (GDH) (EC 1.4.1.2) fromLaccaria bicolorwas purified 410-fold to apparent electrophoretic homogeneity with a 40% recovery through a three-step procedure involving ammonium sulfate precipitation, anion-exchange chromatography on DEAE–Trisacryl, and gel filtration. The molecular weight of the native enzyme determined by gel filtration was 470 kDa, whereas sodium dodecyl sulfate–polyacrylamide gel electrophoresis gave rise to a single band of 116 kDa, suggesting that the enzyme is composed of four identical subunits. The enzyme was specific for NAD(H). The pH optima were 7.4 and 8.8 for the amination and deamination reactions, respectively. The enzyme was found to be highly unstable, with virtually no activity after 20 days at −75°C, 4 days at 4°C, and 1 h at 50°C. The addition of ammonium sulfate improved greatly the stability of the enzyme and full activity was still observed after several months at −75°C. NAD-GDH activity was stimulated by Ca2+and Mg2+but strongly inhibited by Cu2+and slightly by the nucleotides AMP, ADP, and ATP. The Michaelis constants for NAD, NADH, 2-oxoglutarate, and ammonium were 282 μM, 89 μM, 1.35 mM, and 37 mM, respectively. The enzyme had a negative cooperativity for glutamate (Hill number of 0.3), and itsKmvalue increased from 0.24 to 3.6 mM when the glutamate concentration exceeded 1 mM. These affinity constants of the substrates, compared with those of the NADP-GDH of the fungus, suggest that the NAD-GDH is mainly involved in the catabolism of glutamate, while the NADP-GDH is involved in the catalysis of this amino acid.  相似文献   

14.
Epinephrine (EPI) is thought to act by stimulating adenylyl cyclase (ACase) and cAMP production through β-adrenoceptors in the liver of more primitive vertebrates. Recent observations, however, point to an involvement of α1-adrenoceptors in EPI action, at least in some fish species. The role of the α1- and β-adrenergic transduction pathways has been investigated in rainbow trout (Oncorhynchus mykiss) hepatic tissue. Radioligand-binding assays with the β-adrenergic antagonist 3H-CGP-12177 using hepatic membranes purified on a discontinuous sucrose gradient confirmed the presence of β-adrenoceptors (Kd0.36 nM, Bmax 8.61 fmol · mg−1 protein). We provide the first demonstration of α1-adrenoceptors in these same membranes; analysis of binding data with the α1-adrenergic antagonist 3H-prazosin demonstrated a single class of binding sites with a Kdof 15.4 nM and a Bmax of 75.2 fmol · mg−1 protein. There is a straight correlation between β-adrenoceptor occupancy, ACase activation and cAMP production. On the contrary, the role of inositol 1,4,5-trisphosphate (IP3) has to be elucidated; in fact, despite the presence of specific microsomal binding sites for IP3 (Kd 6.03 nM, Bmax 90.2 fmol · mg−1 protein), its cytosolic concentration was not modulated by EPI. On the other hand, we have previously shown in American eel and bullhead hepatocytes that α1-adrenergic agonists are able to increase intracellular concentrations of IP3 and Ca2+ and to activate glycogenolysis. These data suggest a marked variation in the liver of different fish both in terms of α1-binding sites affinity and of α1-adrenoceptor/IP3/Ca2+ transduction systems.  相似文献   

15.
In this study we investigated the ability of Chara intermedia to acclimate to different irradiances (i.e. “low-light” (LL): 20–30 μmol photons m−2 s−1 and “high-light” (HL): 180–200 μmol photons m−2 s−1) and light qualities (white, yellow and green), using morphological, photosynthesis, chlorophyll fluorescence and pigment analysis.Relative growth rates increased with increasing irradiance from 0.016 ± 0.003 (LL) to 0.024 ± 0.005 (HL) g g−1 d−1 fresh weight and were independent of light quality. A growth-based branch orientation towards high-light functioning as a mechanism to protect the plant from excessive light was confirmed. It was shown that the receptor responsible for the morphological reaction is sensitive to blue-light.C. intermedia showed higher oxygen evolution (up to 10.5 (HL) vs. 4.5 (LL) nmol O2 mg Chl−1 s−1), photochemical and energy-dependent Chl fluorescence quenching and a lower Fv/Fm after acclimation to HL. With respect to qP, the acclimation of the photosynthetic apparatus depended on light quality and needed the blue part of the spectrum for full development. In addition, pigment composition was influenced by light and the Chl a/Car and Antheraxanthin (A) + Zeaxanthin (Z)/Violaxanthin (V) + A + Z (DES) ratios revealed the expected acclimation behaviour in favour of carotenoid protection under HL (i.e. decrease of Chl a/Car from 3.41 ± 0.48 to 2.30 ± 0.35 and increase of DES from 0.39 ± 0.05 to 0.87 ± 0.03), while the Chl a/Chl b ratios were not significantly affected. Furthermore it was shown that morphological light acclimation mechanisms influence the extent of the physiological modifications.  相似文献   

16.
An N-acetyl-β-d-hexosaminidase has been purified from primary wheat leaves (Triticum aestivum L.) by freeze-thawing, (NH4)2SO4 precipitation, methanol precipitation, gel filtration, cation exchange chromatography and affinity chromatography on concanavalin A-Sepharose. The activity of the purified preparations could be stabilised by addition of Triton X-100 and the enzyme was stored at -20°C without significant loss of activity. The enzyme hydrolysed pNP-β-d-GlcNAc (optimum pH 5.2, Km 0.29 mM, Vmax 2.56 μkat mg−1) and pNP-β-d-GalNAc (optimum pH 4.4, Km 0.27 mM, Vmax 2.50 μkat mg−1). Five major isozymes were identified, with isoelectric points in the range 5.13–5.36. All five isozymes possessed both N-acety-β-d-glucosaminidase and N-acetyl-β-d-galactosaminidase activity. Inhibition studies and mixed substrate analysis suggested that both substrates are catalysed by the same active site. Both activities were inhibited by GlcNAc, 2-acetamido-2-deoxygluconolactone, GalNAc and the ions of mercury, silver and copper. The Kis for inhibition of N-acetyl-β-d-glucosaminidase activity were: GlcNAc (15.3 mM) and GalNAc (3.4mM). For inhibition of N-acety-β-d-galactosaminidase activity the corresponding values were: GlcNAc (18.2 mM) and GalNac (2.5 mM). The enzyme was considerably less active at releasing pNP from pNP-β-d-(GlcNAc)2 and pNP-β-d-(GlcNAc)3 than from pNP-β-d-GlcNAc. The ability of the N-acetyl-β-d-hexosaminidase to relase GlcNAc from chitin oligomers (GlcNAc)2 (optimum pH 5.0) and (GlcNAc)3−6 (optimum pH 4.4) was also low. Analysis of the reaction products revealed that the initial products from the hydrolysis of (GlcNAc)n were predominantly (GlcNAc)n−1 and GlcNAc.  相似文献   

17.
The cytoplasmic concentrations of Cl([Cl]i) and Ca2+ ([Ca2+]i) were measured with the fluorescent indicators N-(ethoxycarbonylmethyl)-6-methoxyquinilinum bromide (MQAE) and fura-2 in pancreatic β-cells isolated from ob/ob mice. Steady-state [Cl]i in unstimulated β-cells was 34 mM, which is higher than expected from a passive distribution. Increase of the glucose concentration from 3 to 20 mM resulted in an accelerated entry of Cl into β-cells depleted of this ion. The exposure to 20 mM glucose did not affect steady-state [Cl]i either in the absence or presence of furosemide inhibition of Na+, K+, 2 Cl co-transport. Glucose-induced oscillations of [Ca2+]i were transformed into sustained elevation in the presence of 4,4′ diisothiocyanato-dihydrostilbene-2,2′-disulfonic acid (H2DIDS). A similar effect was noted when replacing 25% of extracellular Cl with the more easily permeating anions SCN, I, NO3 or Br. It is concluded that glucose stimulation of the β-cells is coupled to an increase in their Cl permeability and that the oscillatory Ca2+ signalling is critically dependent on transmembrane Cl fluxes.  相似文献   

18.
The heart rate response to isoproterenol (HR-Iso), density and affinity (kd) of β-adrenergic (β-AR) and muscarinic (M2) receptors were compared among three rodents with different generation-life histories of confinement and of high altitude exposure. The European guinea pig (Cavia porcellus) (EGp), a laboratory animal that arrived in Europe after the Spanish Conquest of South America and the Peruvian guinea pig (C. porcellus) (PGp), a semi-wild animal that came from the altiplano to sea level at least 25 generations ago, were used for intra-species comparison. Wistar rats (WR) were used for inter-species comparison as representative of a typical sea level laboratory animal. The HR-Iso was lower in EGp than in the PGp. The PGp showed the highest β-AR density (P<0.0005) and the highest β-AR kd values (P<0.0005) when compared to both EGp and WR groups (β-AR Bmax (fmol mg−1 prot), WR, 19±4; Egp, 34±10; PGp, 74±15. β-AR kd (pM), WR, 24±10; Egp, 17±7; PGp, 39±14). In contrast, PGp showed lower M2 receptor density values than the EGp (P<0.0005). The WR had the highest M2 receptor densities (M2 Bmax (fmol mg−1 prot), WR, 188±15; Egp, 147±9; PGp, 118±6 and M2 kd (pM), WR, 65±12; Egp, 67±6; PGp, 92±2). The inter and intra-species differences found may be related to their respective history of confinement rather than to their history of exposure to high altitude.  相似文献   

19.
Fluorescence redistribution after photobleaching (FRAP) was utilized to select a “fast” lateral mobility clone from Kirsten murine sarcoma virus-transformed 3T3 (KMSV-3T3) fibroblasts. The clone, E7G1, demonstrated a lateral mobility for membrane wheat germ agglutinin (WGA) and succinylated concanavalin A (sCon A) receptors of (2.1 ± 1.6) × 10−9 cm2/s and (2.7 ± 2.3) × 10−9 cm2/s, respectively. These mobilities were approximately equivalent to phospholipid mobility (2.8 ± 1.9 × 10−9 cm2/s). The fast mobility phenotype is observed when the cells are unattached and spherical. Upon attachment, the mobility decreases to (0.19 ± 0.19) × 10−10 cm2/s. In addition, the ability of Con A to initiate global modulation was completely lost in spread as well as spherical cells in the E7G1 fast mobility clone. A comparison of F-actin patterns between untransformed Balb/c fibroblasts and the E7G1-transformed line suggests a correlation between well-developed stress fiber assemblies and the ability to induce global modulation. The fast mobility clone was stable for at least 23 passages.  相似文献   

20.
A 1.4-kb gene encoding the “small” sialidase isoenzyme ofClostridium perfringensA99, including its own promoter, was previously cloned in and expressed byEscherichia coliJM 101. Since all attempts to purify this enzyme to homogeneity were unsuccessful, a new strategy was developed. The structural gene was amplified by means of a PCR technique and inserted into the plasmid vector pQE-10, transferring a six-histidine affinity tag (His6) to the N-terminus of the protein. In order to minimize proteolytic degradation of the sialidase protein, the gene was subcloned into theEscherichia colistrain BL21(DE3)pLys S with reduced protease activity. The sialidase production was increased about 2.5-fold when compared with that of the original clone. The enzyme, released by lysozyme treatment of the bacterial cells, was purified by metal chelate chromatography on Ni–nitrilo-triacetic acid agarose to apparent homogeneity in SDS–PAGE. The 42-kDa protein was enriched 62-fold with a yield of 82% and a specific activity of 280 U mg−1. A total amount of 1 mg sialidase was obtained from 1 liter of bacterial culture. For future studies, including crystallization experiments, the histidine affinity tag was removed from the sialidase enzyme by aminopeptidase K. The sialidase was then separated from aminopeptidase K by ion-exchange chromatography, resulting in an overall yield of 83% and a specific activity of 305 U mg−1using 4-methylumbelliferyl-α- -N-acetylneuraminic acid under standard conditions. The two forms (with or without the histidine tag) of sialidase exhibited similar kinetic properties when compared to the wild-type enzyme.  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