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 359 毫秒
1.
We purified a secreted fungal laccase from filtrates of Gaeumannomyces graminis var. tritici cultures induced with copper and xylidine. The active protein had an apparent molecular mass of 190 kDa and yielded subunits with molecular masses of 60 kDa when denatured and deglycosylated. This laccase had a pI of 5.6 and an optimal pH of 4.5 with 2,6-dimethoxyphenol as its substrate. Like other, previously purified laccases, this one contained several copper atoms in each subunit, as determined by inductively coupled plasma spectroscopy. The active enzyme catalyzed the oxidation of 2,6-dimethoxyphenol (Km = 2.6 × 10−5 ± 7 × 10−6 M), catechol (Km = 2.5 × 10−4 ± 1 × 10−5 M), pyrogallol (Km = 3.1 × 10−4 ± 4 × 10−5 M), and guaiacol (Km = 5.1 × 10−4 ± 2 × 10−5 M). In addition, the laccase catalyzed the polymerization of 1,8-dihydroxynaphthalene, a natural fungal melanin precursor, into a high-molecular-weight melanin and catalyzed the oxidation, or decolorization, of the dye poly B-411, a lignin-like polymer. These findings indicate that this laccase may be involved in melanin polymerization in this phytopathogen’s hyphae and/or in lignin depolymerization in its infected plant host.  相似文献   

2.
Myrosinase (β-thioglucoside glucohydrolase, EC 3.2.3.1) was purified to apparent homogeneity from light-grown cress (Lepidium sativum L.) seedlings. This enzyme, which catalyzes hydrolysis of the glucosinolate sinigrin (Km, 115 micromolar) at an optimum pH of 5.5 in sodium citrate buffer, had a native molecular weight of 130 ± 5 kilodaltons and an isoelectric point of 4.7 to 4.9. SDS-PAGE revealed two polypeptides with molecular weights of 62 and 65 kilodaltons. Both subunits contained carbohydrate as shown by periodic acid-Schiff staining. The purified enzyme hydrolyzed p-nitrophenyl-β-d-glucoside (Km, 2.0 millimolar) at an optimum pH of 6.5 in phosphate buffer. The indolizidine alkaloid castanospermine, a known inhibitor of O-glycosidases, competitively inhibited the hydrolyses of sinigrin (thioglucosidase activity) and p-nitrophenyl-β-d-glucoside (O-glucosidase activity) with Ki values of 5 and 6 micromolar, respectively. In contrast, the related polyhydroxyalkaloids swainsonine and deoxynojirimycin were without effect upon these hydrolyses.  相似文献   

3.
Two Salmonella typhimurium strains, which could be used as sources for the leucine biosynthetic intermediates α- and β-isopropylmalate were constructed by a series of P22-mediated transductions. One strain, JK527 [flr-19 leuA2010 Δ(leuD-ara)798 fol-162], accumulated and excreted α-isopropylmalate, whereas the second strain, JK553 (flr-19 leuA2010 leuB698), accumulated and excreted α- and β-isopropylmalate. The yield of α-isopropylmalate isolated from the culture medium of JK527 was more than five times the amount obtained from a comparable volume of medium in which Neurospora crassa strain FLR92-1-216 (normally used as the source for α- and β-isopropylmalate) was grown. Not only was the yield greater, but S. typhimurium strains are much easier to handle and grow to saturation much faster than N. crassa strains. The combination of the two regulatory mutations flr-19, which results in constitutive expression of the leucine operon, and leuA2010, which renders the first leucine-specific biosynthetic enzyme insensitive to feedback inhibition by leucine, generated limitations in the production of valine and pantothenic acid. The efficient, irreversible, and unregulated conversion of α-ketoisovaleric acid into α-isopropylmalate (α-isopropylmalate synthetase Km for α-ketoisovaleric acid, 6 × 10−5 M) severely restricted the amount of α-ketoisovaleric acid available for conversion into valine and pantothenic acid (ketopantoate hydroxymethyltransferase Km for α-ketoisovaleric acid, 1.1 × 10−3 M; transaminase B Km for α-ketoisovaleric acid, 2 × 10−3 M).  相似文献   

4.
Bacterial steroid catabolism is an important component of the global carbon cycle and has applications in drug synthesis. Pathways for this catabolism involve multiple acyl coenzyme A (CoA) synthetases, which activate alkanoate substituents for β-oxidation. The functions of these synthetases are poorly understood. We enzymatically characterized four distinct acyl-CoA synthetases from the cholate catabolic pathway of Rhodococcus jostii RHA1 and the cholesterol catabolic pathway of Mycobacterium tuberculosis. Phylogenetic analysis of 70 acyl-CoA synthetases predicted to be involved in steroid metabolism revealed that the characterized synthetases each represent an orthologous class with a distinct function in steroid side-chain degradation. The synthetases were specific for the length of alkanoate substituent. FadD19 from M. tuberculosis H37Rv (FadD19Mtb) transformed 3-oxo-4-cholesten-26-oate (kcat/Km = 0.33 × 105 ± 0.03 × 105 M−1 s−1) and represents orthologs that activate the C8 side chain of cholesterol. Both CasGRHA1 and FadD17Mtb are steroid-24-oyl-CoA synthetases. CasG and its orthologs activate the C5 side chain of cholate, while FadD17 and its orthologs appear to activate the C5 side chain of one or more cholesterol metabolites. CasIRHA1 is a steroid-22-oyl-CoA synthetase, representing orthologs that activate metabolites with a C3 side chain, which accumulate during cholate catabolism. CasI had similar apparent specificities for substrates with intact or extensively degraded steroid nuclei, exemplified by 3-oxo-23,24-bisnorchol-4-en-22-oate and 1β(2′-propanoate)-3aα-H-4α(3″-propanoate)-7aβ-methylhexahydro-5-indanone (kcat/Km = 2.4 × 105 ± 0.1 × 105 M−1 s−1 and 3.2 × 105 ± 0.3 × 105 M−1 s−1, respectively). Acyl-CoA synthetase classes involved in cholate catabolism were found in both Actinobacteria and Proteobacteria. Overall, this study provides insight into the physiological roles of acyl-CoA synthetases in steroid catabolism and a phylogenetic classification enabling prediction of specific functions of related enzymes.  相似文献   

5.
Oscillatory behavior of mitochondrial inner membrane potential (ΔΨm) is commonly observed in cells subjected to oxidative or metabolic stress. In cardiac myocytes, the activation of inner membrane pores by reactive oxygen species (ROS) is a major factor mediating intermitochondrial coupling, and ROS-induced ROS release has been shown to underlie propagated waves of ΔΨm depolarization as well as synchronized limit cycle oscillations of ΔΨm in the network. The functional impact of ΔΨm instability on cardiac electrophysiology, Ca2+ handling, and even cell survival, is strongly affected by the extent of such intermitochondrial coupling. Here, we employ a recently developed wavelet-based analytical approach to examine how different substrates affect mitochondrial coupling in cardiac cells, and we also determine the oscillatory coupling properties of mitochondria in ventricular cells in intact perfused hearts. The results show that the frequency of ΔΨm oscillations varies inversely with the size of the oscillating mitochondrial cluster, and depends on the strength of local intermitochondrial coupling. Time-varying coupling constants could be quantitatively determined by applying a stochastic phase model based on extension of the well-known Kuramoto model for networks of coupled oscillators. Cluster size-frequency relationships varied with different substrates, as did mitochondrial coupling constants, which were significantly larger for glucose (7.78 × 10−2 ± 0.98 × 10−2 s−1) and pyruvate (7.49 × 10−2 ± 1.65 × 10−2 s−1) than lactate (4.83 × 10−2 ± 1.25 × 10−2 s−1) or β-hydroxybutyrate (4.11 × 10−2 ± 0.62 × 10−2 s−1). The findings indicate that mitochondrial spatiotemporal coupling and oscillatory behavior is influenced by substrate selection, perhaps through differing effects on ROS/redox balance. In particular, glucose-perfusion generates strong intermitochondrial coupling and temporal oscillatory stability. Pathological changes in specific catabolic pathways, which are known to occur during the progression of cardiovascular disease, could therefore contribute to altered sensitivity of the mitochondrial network to oxidative stress and emergent ΔΨm instability, ultimately scaling to produce organ level dysfunction.  相似文献   

6.
Whole-cell assays of methane and trichloroethylene (TCE) consumption have been performed on Methylosinus trichosporium OB3b expressing particulate methane monooxygenase (pMMO). From these assays it is apparent that varying the growth concentration of copper causes a change in the kinetics of methane and TCE degradation. For M. trichosporium OB3b, increasing the copper growth concentration from 2.5 to 20 μM caused the maximal degradation rate of methane (Vmax) to decrease from 300 to 82 nmol of methane/min/mg of protein. The methane concentration at half the maximal degradation rate (Ks) also decreased from 62 to 8.3 μM. The pseudo-first-order rate constant for methane, Vmax/Ks, doubled from 4.9 × 10−3 to 9.9 × 10−3 liters/min/mg of protein, however, as the growth concentration of copper increased from 2.5 to 20 μM. TCE degradation by M. trichosporium OB3b was also examined with varying copper and formate concentrations. M. trichosporium OB3b grown with 2.5 μM copper was unable to degrade TCE in both the absence and presence of an exogenous source of reducing equivalents in the form of formate. Cells grown with 20 μM copper, however, were able to degrade TCE regardless of whether formate was provided. Without formate the Vmax for TCE was 2.5 nmol/min/mg of protein, while providing formate increased the Vmax to 4.1 nmol/min/mg of protein. The affinity for TCE also increased with increasing copper, as seen by a change in Ks from 36 to 7.9 μM. Vmax/Ks for TCE degradation by pMMO also increased from 6.9 × 10−5 to 5.2 × 10−4 liters/min/mg of protein with the addition of formate. From these whole-cell studies it is apparent that the amount of copper available is critical in determining the oxidation of substrates in methanotrophs that are expressing only pMMO.  相似文献   

7.
1. Purified rabbit-muscle and -liver glucose phosphate isomerase, free of contaminating enzyme activities that could interfere with the assay procedures, were tested for inhibition by fructose, fructose 1-phosphate and fructose 1,6-diphosphate. 2. Fructose 1-phosphate and fructose 1,6-diphosphate are both competitive with fructose 6-phosphate in the enzymic reaction, the apparent Ki values being 1·37×10−3−1·67×10−3m for fructose 1-phosphate and 7·2×10−3−7·9×10−3m for fructose 1,6-diphosphate; fructose and inorganic phosphate were without effect. 3. The apparent Km values for both liver and muscle enzymes at pH7·4 and 30° were 1·11×10−4−1·29×10−4m for fructose 6-phosphate, determined under the conditions in this paper. 4. In the reverse reaction, fructose, fructose 1-phosphate and fructose 1,6-diphosphate did not significantly inhibit the conversion of glucose 6-phosphate into fructose 6-phosphate. 5. The apparent Km values for glucose 6-phosphate were in the range 5·6×10−4−8·5×10−4m. 6. The competitive inhibition of hepatic glucose phosphate isomerase by fructose 1-phosphate is discussed in relation to the mechanism of fructose-induced hypoglycaemia in hereditary fructose intolerance.  相似文献   

8.
Marine microbes use alginate lyases to degrade and catabolize alginate, a major cell wall matrix polysaccharide of brown seaweeds. Microbes frequently contain multiple, apparently redundant alginate lyases, raising the question of whether these enzymes have complementary functions. We report here on the molecular cloning and functional characterization of three exo-type oligoalginate lyases (OalA, OalB, and OalC) from Vibrio splendidus 12B01 (12B01), a marine bacterioplankton species. OalA was most active at 16°C, had a pH optimum of 6.5, and displayed activities toward poly-β-d-mannuronate [poly(M)] and poly-α-l-guluronate [poly(G)], indicating that it is a bifunctional enzyme. OalB and OalC were most active at 30 and 35°C, had pH optima of 7.0 and 7.5, and degraded poly(M·G) and poly(M), respectively. Detailed kinetic analyses of oligoalginate lyases with poly(G), poly(M), and poly(M·G) and sodium alginate as substrates demonstrated that OalA and OalC preferred poly(M), whereas OalB preferred poly(M·G). The catalytic efficiency (kcat/Km) of OalA against poly(M) increased with decreasing size of the substrate. OalA showed kcat/Km from 2,130 mg−1 ml s−1 for the trisaccharide to 224 mg−1 ml s−1 for larger oligomers of ∼50 residues, and 50.5 mg−1 ml s−1 for high-molecular-weight alginate. Although OalA was most active on the trisaccharide, OalB and OalC preferred dimers. Taken together, our results indicate that these three Oals have complementary substrate scopes and temperature and pH adaptations.  相似文献   

9.
Initial reactions involved in the bacterial degradation of polycyclic aromatic hydrocarbons (PAHs) include a ring-dihydroxylation catalyzed by a dioxygenase and a subsequent oxidation of the dihydrodiol products by a dehydrogenase. In this study, the dihydrodiol dehydrogenase from the PAH-degrading Sphingomonas strain CHY-1 has been characterized. The bphB gene encoding PAH dihydrodiol dehydrogenase (PDDH) was cloned and overexpressed as a His-tagged protein. The recombinant protein was purified as a homotetramer with an apparent Mr of 110,000. PDDH oxidized the cis-dihydrodiols derived from biphenyl and eight polycyclic hydrocarbons, including chrysene, benz[a]anthracene, and benzo[a]pyene, to corresponding catechols. Remarkably, the enzyme oxidized pyrene 4,5-dihydrodiol, whereas pyrene is not metabolized by strain CHY-1. The PAH catechols produced by PDDH rapidly auto-oxidized in air but were regenerated upon reaction of the o-quinones formed with NADH. Kinetic analyses performed under anoxic conditions revealed that the enzyme efficiently utilized two- to four-ring dihydrodiols, with Km values in the range of 1.4 to 7.1 μM, and exhibited a much higher Michaelis constant for NAD+ (Km of 160 μM). At pH 7.0, the specificity constant ranged from (1.3 ± 0.1) × 106 M−1 s−1 with benz[a]anthracene 1,2-dihydrodiol to (20.0 ± 0.8) × 106 M−1 s−1 with naphthalene 1,2-dihydrodiol. The catalytic activity of the enzyme was 13-fold higher at pH 9.5. PDDH was subjected to inhibition by NADH and by 3,4-dihydroxyphenanthrene, and the inhibition patterns suggested that the mechanism of the reaction was ordered Bi Bi. The regulation of PDDH activity appears as a means to prevent the accumulation of PAH catechols in bacterial cells.  相似文献   

10.
Using a combined experimental and theoretical approach named binding-unbinding correlation spectroscopy (BUCS), we describe the two-dimensional kinetics of interactions between fibrinogen and the integrin αIIbβ3, the ligand-receptor pair essential for platelet function during hemostasis and thrombosis. The methodology uses the optical trap to probe force-free association of individual surface-attached fibrinogen and αIIbβ3 molecules and forced dissociation of an αIIbβ3-fibrinogen complex. This novel approach combines force clamp measurements of bond lifetimes with the binding mode to quantify the dependence of the binding probability on the interaction time. We found that fibrinogen-reactive αIIbβ3 pre-exists in at least two states that differ in their zero force on-rates (kon1 = 1.4 × 10−4 and kon2 = 2.3 × 10−4 μm2/s), off-rates (koff1 = 2.42 and koff2 = 0.60 s−1), and dissociation constants (Kd1 = 1.7 × 104 and Kd2 = 2.6 × 103 μm−2). The integrin activator Mn2+ changed the on-rates and affinities (Kd1 = 5 × 104 and Kd2 = 0.3 × 103 μm−2) but did not affect the off-rates. The strength of αIIbβ3-fibrinogen interactions was time-dependent due to a progressive increase in the fraction of the high affinity state of the αIIbβ3-fibrinogen complex characterized by a faster on-rate. Upon Mn2+-induced integrin activation, the force-dependent off-rates decrease while the complex undergoes a conformational transition from a lower to higher affinity state. The results obtained provide quantitative estimates of the two-dimensional kinetic rates for the low and high affinity αIIbβ3 and fibrinogen interactions at the single molecule level and offer direct evidence for the time- and force-dependent changes in αIIbβ3 conformation and ligand binding activity, underlying the dynamics of fibrinogen-mediated platelet adhesion and aggregation.  相似文献   

11.
Enzymes in cancer: Asparaginase from chicken liver   总被引:2,自引:1,他引:1  
1. A procedure for partial purification of asparaginase from chicken liver is presented. 2. The bulk of the enzyme is located in the soluble fraction of chicken liver. 3. Molecular weights of chicken-liver asparaginase and of the guinea-pig serum enzyme, estimated by gel filtration, were 306000 and 210000 respectively. The Michaelis constants (Km) at 37° and pH8·5 were 6·0×10−5m and 7·2×10−5m respectively. 4. At 50° the chicken-liver enzyme was moderately stable, some activity being lost by aggregation; in dilute electrolyte solutions the activity rapidly diminished. 5. The anti-lymphoma effect of guinea-pig serum in mice carrying the 6C3HED tumour was confirmed. Chicken-liver asparaginase also showed an effect but in this case the enzyme preparation had to be administered repeatedly. 6. Guinea-pig serum asparaginase was stable for several days in mouse blood, after intraperitoneal injection, whereas chicken-liver asparaginase rapidly disappeared. 7. Aspartic acid β-hydrazide was shown to be a competitive inhibitor of chicken-liver asparaginase with Ki approx. 5·6×10−4m. In mice it produced an anti-lymphoma effect, as reported previously.  相似文献   

12.
Avibactam is a non-β-lactam β-lactamase inhibitor with a spectrum of activity that includes β-lactamase enzymes of classes A, C, and selected D examples. In this work acylation and deacylation rates were measured against the clinically important enzymes CTX-M-15, KPC-2, Enterobacter cloacae AmpC, Pseudomonas aeruginosa AmpC, OXA-10, and OXA-48. The efficiency of acylation (k2/Ki) varied across the enzyme spectrum, from 1.1 × 101 m−1s−1 for OXA-10 to 1.0 × 105 for CTX-M-15. Inhibition of OXA-10 was shown to follow the covalent reversible mechanism, and the acylated OXA-10 displayed the longest residence time for deacylation, with a half-life of greater than 5 days. Across multiple enzymes, acyl enzyme stability was assessed by mass spectrometry. These inhibited enzyme forms were stable to rearrangement or hydrolysis, with the exception of KPC-2. KPC-2 displayed a slow hydrolytic route that involved fragmentation of the acyl-avibactam complex. The identity of released degradation products was investigated, and a possible mechanism for the slow deacylation from KPC-2 is proposed.  相似文献   

13.
Isolation and Characterization of a Xylanase from Bacillus subtilis   总被引:4,自引:3,他引:1       下载免费PDF全文
Partial characterization of an extracellular xylanase isolated by chromatography from Bacillus subtilis gave a molecular weight of 32,000 and optimum pH and temperature of 5.0 and 50°C, respectively. Km and Vmax values, determined with a soluble larchwood xylan, were 0.16% and 7.0 × 103 μmol min−1 mg−1 of enzyme respectively. The amino acid composition showed more basic amino acid residues than in a previously characterized xylanase from a white-rot fungus.  相似文献   

14.
A new oriented method using a diazonium salt reaction was developed for linking β 2-adrenoceptor (β 2-AR) on the surface of macroporous silica gel. Stationary phase containing the immobilised receptor was used to investigate the interaction between β 2-AR and ephedrine plus pseudoephedrine by zonal elution. The isotherms of the two drugs best fit the Langmuir model. Only one type of binding site was found for ephedrine and pseudoephedrine targeting β 2-AR. At 37 °C, the association constants during the binding were (5.94±0.05)×103/M for ephedrine and (3.80±0.02) ×103/M for pseudoephedrine, with the binding sites of (8.92±0.06) ×10−4 M. Thermodynamic studies showed that the binding of the two compounds to β 2-AR was a spontaneous reaction with exothermal processes. The ΔGθ, ΔHθ and ΔSθ for the interaction between ephedrine and β 2-AR were −(22.33±0.04) kJ/mol, −(6.51±0.69) kJ/mol and 50.94±0.31 J/mol·K, respectively. For the binding of pseudoephedrine to the receptor, these values were −(21.17±0.02) kJ/mol, −(7.48±0.56) kJ/mol and 44.13±0.01 J/mol·K. Electrostatic interaction proved to be the driving force during the binding of the two drugs to β 2-AR. The proposed immobilised method will have great potential for attaching protein to solid substrates and realizing the interactions between proteins and drugs.  相似文献   

15.
Ceramide glucoside (1-O-glucosido-2-N-acyl-sphingosine) was hydrolysed to ceramide (N-acyl-sphingosine) and glucose by β-glucosidase from ox brain. The reaction was stimulated by the non-ionic detergent, Triton X-100, or by the anionic detergents, cholate or taurocholate. It was not reversible, had optimum pH5·0 (with acetate buffer) or 5·6 (with pyridine buffer), had Km 1·8×10−4m and was inhibited by δ-gluconolactone and sphingosine, but not by ceramide or palmitic acid.  相似文献   

16.
Burkholderia territorii, a Gram-negative bacterium, encodes for the ι-class carbonic anhydrase (CA, EC 4.2.1.1) BteCAι, which was recently characterised. It acts as a good catalyst for the hydration of CO2 to bicarbonate and protons, with a kcat value of 3.0 × 105 s−1 and kcat/KM value of 3.9 × 107 M−1 s−1. No inhibition data on this new class of enzymes are available to date. We report here an anion and small molecules inhibition study of BteCAι, which we prove to be a zinc(II)- and not manganese(II)-containing enzyme, as reported for diatom ι-CAs. The best inhibitors were sulphamic acid, stannate, phenylarsonic acid, phenylboronic acid and sulfamide (KI values of 6.2–94 µM), whereas diethyldithiocarbamate, tellurate, selenate, bicarbonate and cyanate were submillimolar inhibitors (KI values of 0.71–0.94 mM). The halides (except iodide), thiocyanate, nitrite, nitrate, carbonate, bisulphite, sulphate, hydrogensulfide, peroxydisulfate, selenocyanate, fluorosulfonate and trithiocarbonate showed KI values in the range of 3.1–9.3 mM.  相似文献   

17.
An alkaliphilic, thermophilic Bacillus sp. (NCIM 59) produced extracellular xylose isomerase at pH 10 and 50°C by using xylose or wheat bran as the carbon source. The distribution of xylose isomerase as a function of growth in comparison with distributions of extra- and intracellular marker enzymes such as xylanase and β-galactosidase revealed that xylose isomerase was truly secreted as an extracellular enzyme and was not released because of sporulation or lysis. The enzyme was purified to homogeneity by ammonium sulfate precipitation followed by gel filtration, preparative polyacrylamide gel electrophoresis, and ion-exchange chromatography. The molecular weight of xylose isomerase was estimated to be 160,000 by gel filtration and 50,000 by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, indicating the presence of three subunits. The enzyme is most active at pH 8.0 and with incubation at 85°C for 20 min. Divalent metal ions Mg2+, Co2+, and Mn2+ were required for maximum activity of the enzyme. The Km values for D-xylose and D-glucose at 80°C and pH 7.5 were 6.66 and 142 mM, respectively, while Kcat values were 2.3 × 102 s-1 and 0.5 × 102 s-1, respectively.  相似文献   

18.
Current-voltage curves for DIDS-insensitive Cl conductance have been determined in human red blood cells from five donors. Currents were estimated from the rate of cell shrinkage using flow cytometry and differential laser light scattering. Membrane potentials were estimated from the extracellular pH of unbuffered suspensions using the proton ionophore FCCP. The width of the Gaussian distribution of cell volumes remained invariant during cell shrinkage, indicating a homogeneous Cl conductance among the cells. After pretreatment for 30 min with DIDS, net effluxes of K+ and Cl were induced by valinomycin and were measured in the continued presence of DIDS; inhibition was maximal at ∼65% above 1 μM DIDS at both 25°C and 37°C. The nonlinear current-voltage curves for DIDS-insensitive net Cl effluxes, induced by valinomycin or gramicidin at varied [K+]o, were compared with predictions based on (1) the theory of electrodiffusion, (2) a single barrier model, (3) single occupancy, multiple barrier models, and (4) a voltage-gated mechanism. Electrodiffusion precisely describes the relationship between the measured transmembrane voltage and [K+]o. Under our experimental conditions (pH 7.5, 23°C, 1–3 μM valinomycin or 60 ng/ml gramicidin, 1.2% hematocrit), the constant field permeability ratio PK/PCl is 74 ± 9 with 10 μM DIDS, corresponding to 73% inhibition of PCl. Fitting the constant field current-voltage equation to the measured Cl currents yields P Cl = 0.13 h−1 with DIDS, compared to 0.49 h−1 without DIDS, in good agreement with most previous studies. The inward rectifying DIDS-insensitive Cl current, however, is inconsistent with electrodiffusion and with certain single-occupancy multiple barrier models. The data are well described either by a single barrier located near the center of the transmembrane electric field, or, alternatively, by a voltage-gated channel mechanism according to which the maximal conductance is 0.055 ± 0.005 S/g Hb, half the channels are open at −27 ± 2 mV, and the equivalent gating charge is −1.2 ± 0.3.  相似文献   

19.
Synthetic and natural polymers are often used as drug delivery systems in vitro and in vivo. Biodegradable chitosan of different sizes were used to encapsulate antitumor drug tamoxifen (Tam) and its metabolites 4-hydroxytamoxifen (4-Hydroxytam) and endoxifen (Endox). The interactions of tamoxifen and its metabolites with chitosan 15, 100 and 200 KD were investigated in aqueous solution, using FTIR, fluorescence spectroscopic methods and molecular modeling. The structural analysis showed that tamoxifen and its metabolites bind chitosan via both hydrophilic and hydrophobic contacts with overall binding constants of K tam-ch-15  = 8.7 (±0.5)×103 M−1, K tam-ch-100  = 5.9 (±0.4)×105 M−1, K tam-ch-200  = 2.4 (±0.4)×105 M−1 and K hydroxytam-ch-15  = 2.6(±0.3)×104 M−1, K hydroxytam – ch-100  = 5.2 (±0.7)×106 M−1 and K hydroxytam-ch-200  = 5.1 (±0.5)×105 M−1, K endox-ch-15  = 4.1 (±0.4)×103 M−1, K endox-ch-100  = 1.2 (±0.3)×106 M−1 and K endox-ch-200  = 4.7 (±0.5)×105 M−1 with the number of drug molecules bound per chitosan (n) 2.8 to 0.5. The order of binding is ch-100>200>15 KD with stronger complexes formed with 4-hydroxytamoxifen than tamoxifen and endoxifen. The molecular modeling showed the participation of polymer charged NH2 residues with drug OH and NH2 groups in the drug-polymer adducts. The free binding energies of −3.46 kcal/mol for tamoxifen, −3.54 kcal/mol for 4-hydroxytamoxifen and −3.47 kcal/mol for endoxifen were estimated for these drug-polymer complexes. The results show chitosan 100 KD is stronger carrier for drug delivery than chitosan-15 and chitosan-200 KD.  相似文献   

20.
α-Amylase (EC 3.2.1.1) was excreted by Calvatia gigantea in liquid growth media containing different sources of starch. Among the factors affecting enzyme production in shake flasks were the type and the concentration of starch and the nitrogen source supplied. Optimum cultural conditions for maximum enzyme production were: soluble starch concentration, 5%; inoculum size, 3.75 × 105 conidia per ml; 5-day cultivation time at 28 to 30°C. The observed maximum yield of 81.3 U of saccharifying enzyme activity per ml of growth medium was the highest ever reported in the literature for submerged cultures. Partially purified enzyme functioned optimally at pH 4.5 to 5.5 and 53 to 58°C. The activation energy of enzymic hydrolysis of starch in the range of 20 to 40°C was 8,125 cal/mol (ca. 3.41 × 104 J). The apparent Km value of the enzyme at 25°C was 7.68 × 10−4 g/ml. Some of the properties of the enzyme under investigation were similar to those of α-amylases excreted from molds producing large amounts of the 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