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 93 毫秒
1.
Extraction of control human spleen glucocerebrosidase with sodium cholate and butan-l-ol reversibly inactivates the enzyme in terms of its ability to hydrolyse the water-soluble substrate 4-methylumbelliferyl beta-D-glucopyranoside (MUGlc). The acidic brain lipid galactocerebroside 3-sulphate (sulphatide) reconstitutes beta-glucosidase activity in a strongly concentration-dependent manner. In this study we show that sulphatide exhibits three critical micellar concentrations (CMCs): CMC1, 3.72 microM; CMC2, 22.6 microM; CMC3, 60.7 microM. We designate the aggregates formed at these CMCs as primary, secondary and tertiary micelles respectively. From the results of kinetic studies performed at various sulphatide concentrations (0.012-248 microM), we found that sulphatide monomers (less than 3 microM) decreased the Km (for MUGlc) of control glucocerebrosidase from 11 to 4.6 mM, and lowered the Vmax. 2-fold. However, secondary and tertiary micelles were required for expression of high control glucocerebrosidase activities. Glucocerebrosidase prepared from the spleen of a patient with non-neuronopathic type 1 Gaucher's disease exhibited a very low Km (2.8 mM) even in the absence of exogenous lipid, and sulphatide monomers had no effect on the mutant enzyme's Km or Vmax. However, secondary or tertiary micelles markedly increased the Vmax. of the type 1 glucocerebrosidase to 60% of the corresponding control enzyme value. In contrast, for the glucocerebrosidase of the neuronopathic type 2 case, although sulphatide decreased the Km from 9.2 to 1.7 mM, the Vmax. never reached more than 5% that of the control enzyme, even at high concentrations of sulphatide. In addition, we found that secondary and tertiary sulphatide micelles enhanced the rate of inactivation of all three glucocerebrosidase preparations by chymotrypsin. Collectively, these results indicate the presence of two sulphatide-binding sites on glucocerebrosidase: one that enhances substrate binding, and another that enhances catalysis.  相似文献   

2.
The lipid requirement of membrane-bound rat liver beta-glucosidase was investigated using 4-methylumbelliferyl-beta-D-glucopyranoside as the substrate. The enzyme was solubilized and delipidated by sequential extraction of a crude lysosomal fraction from rat liver lysosomes with sodium cholate and ice-cold butan-1-ol. Neither saturated nor unsaturated phosphatidylcholine activated this enzyme. In contrast, acidic phospholipids like phosphatidylglycerol (PtdGro) and phosphatidylserine (PtdSer) were effective activators. For the PtdGro series, fatty acid composition was important, with the shorter chain or unsaturated fatty acid-containing PtdGro species being the best activators. Heat-stable factor (HSF) from Gaucher spleen by itself (1-2 micrograms) had no effect on enzyme activity. However, the same amount of HSF when combined with 10 micrograms of PtdSer markedly stimulated beta-glucosidase activity. In the presence of HSF, di-9-cis-octadecenoyl-PtdGro (1 microgram) or -PtdSer (5 micrograms) provided maximum protection of beta-glucosidase against heat (60 degrees C) inactivation. In the absence of phospholipids, HSF had no effect on the rate of inactivation of the enzyme by the suicide inhibitor conduritol B epoxide (t0.5, 12 +/- 0.5 min); the maximum rate of inactivation was achieved in the presence of a mixture of PtdGro (2.5-5 micrograms) and HSF (t0.5, 2.8 min). The combination of PtdSer (10 micrograms) and HSF (1.3 micrograms) lowered the Km for 4-methylumbelliferyl-beta-D-glucopyranoside from 24 to 2.7 mM. Inhibition of the enzyme by the glucocerebrosidase substrate analogues N-hexyl-O-glucosylsphingosine and glucosylsphingosine was influenced by the activator substances. The inclusion of PtdSer and HSF in the beta-glucosidase assay medium lowered the Ki of N-hexyl-O-glucosylsphingosine 20-fold. The same combination of activators decreased the I0.5 of the enzyme for glucosylsphingosine from 89.4 to 7.6 microM. A study of log (Vmax./Km) versus pH indicated that the PtdSer-HSF pair creates the active site of beta-glucosidase, making apparent three ionizable groups on the enzyme with pK values in the range 4.5-5.1.  相似文献   

3.
Using sucrose density gradients, differences in delipidated lysosomal beta-glucosidase isolated from control spleen and spleen from patients with nonneurologic (type 1) and neurologic (type 2) Gaucher's disease have been examined. The three enzymes differ in sedimentation properties as well as in their responsiveness to activation by phosphatidylserine and heat-stable factor. The control beta-glucosidase sedimented as an apparent 45,000-Da species whose activity was dependent upon the inclusion of exogenous sodium taurodeoxycholate in the assay medium. Preincubation with a mixture of phosphatidylserine and heat-stable factor converted the control enzyme to a faster-sedimenting form which exhibited considerable activity in the absence of exogenous bile salt. Spleen beta-glucosidase from a patient with type 1 Gaucher's disease exhibited an apparent molecular weight of 154,000 on sucrose gradients. Like the control enzyme, the activity of this form was bile salt dependent. Upon preincubation with phosphatidylserine and heat-stable factor, beta-glucosidase from the type 1 case was also converted to a faster-sedimenting form which was more active in the absence of sodium taurodeoxycholate than in the presence of the bile salt. Spleen beta-glucosidase from the patient with type 2 Gaucher's disease sedimented as a broad peak of activity in the most dense regions of the sucrose gradients, appearing to be much larger than the beta-glucosidase from either the control or the type 1 Gaucher's disease patient. The activity of this large species was strongly dependent upon bile salt, and was not affected by preincubation of the enzyme with phosphatidylserine and heat-stable factor. Using the chaotropic salt, sodium thiocyanate (0.15 M), the spleen beta-glucosidase isolated from the type 1 Gaucher's disease case was converted to a slower-sedimenting species. The control enzyme sedimented slightly farther into the sucrose gradients upon treatment with the NaSCN. Thiocyanate treatment had no effect on the spleen beta-glucosidase isolated from the case of type 2 Gaucher's disease.  相似文献   

4.
The micellar properties of mixtures of GM1 ganglioside and the non-ionic amphiphile Triton X-100 in 25 mM Na phosphate-5 mM di Na EDTA buffer (pH = 7.0) were investigated by quasielastic light scattering in a wide range of Triton/GM1 molar ratios and in the temperature range 15–37°C. These measurements: (a) provided evidence for the formation of mixed micelles; (b) allowed the determination of such parameters as the molecular weight and the hydrodynamic radius of the mixed micelles; (c) showed the occurrence of statistical aggregates of micelles with increasing temperature and micelle concentration. Galactose oxidase was chosen for studying the relation between enzyme activity and micellar properties. The action of the enzyme on GM1 was found to be strongly dependent on the micellar structure. In particular: (a) galactose oxidase acted very poorly on homogeneous GM1 micelles, while affecting mixed GM1/Triton X-100 micelles; (b) at fixed GM1 concentration the oxidation rate increased by enhancing Triton X-100 concentration and followed a biphasic kinetics with a break at a certain Triton X-100 concentration; (c) the formation of statistical micelle aggregates was followed by inhibition of the enzyme activity.  相似文献   

5.
This report demonstrates the effect of primary alcohols on the kinetic properties of guinea pig liver cytosolic beta-glucosidase. Lineweaver-Burk analyses of the kinetic data revealed a biphasic response; at low concentrations the alcohols increased the Vmax 5--7-fold while at higher concentrations they caused a purely competitive type of inhibition. For example, with n-butyl alcohol, increasing the alcohol's concentration in the assay medium from 0 to 0.14 M (0-1% (v/v)) resulted in a progressive increase in Vmax to a value 7-fold above the basal level without affecting the Km. However, between 0.14 and 0.54 M (1 and 4% (v/v)) n-butyl alcohol, the Km for 4-methylumbelliferyl-beta-D-glucopyranoside increased significantly from 0.14 to 0.93 mM. In contrast to n-butyl alcohol or isobutyl alcohol, which are potent activators, structurally related compounds like sec-butyl alcohol, tert-butyl alcohol, butylurea, and butanesulfonic acid did not stimulate the activity of the cytosolic beta-glucosidase. In the concentration range where activation was observed, conventional secondary replots of 1/delta slope versus 1/[alcohol] yielded perfect straight lines, demonstrating that binding of a single molecule of alcohol to the beta-glucosidase was responsible for the initial phase of activation. Furthermore, the glycohydrolase displayed a propensity to bind the longer chain alcohols, as reflected by the KA (binding constant) values of 555, 146, 34.1, and 7.47 mM for ethanol, n-propyl alcohol, n-butyl alcohol, 1-pentanol, respectively. This phenomenon of nonessential activation by alcohols has led us to speculate on the presence of a physiologic activator for the beta-glucosidase in mammalian tissues which contain this enzyme.  相似文献   

6.
1. Glucocerebrosidase, in association with a membrane-bound ;acid' beta-glucosidase, was separated from a soluble ;neutral' beta-glucosidase that had no activity towards glucocerebroside as substrate. 2. Glucocerebrosidase, as well as ;acid' beta-glucosidase activity depended upon the association of factor P (a heat-stable, soluble, acidic glycoprotein) with factor C (a heat-labile membrane-bound protein). 3. Factor C was solubilized under certain conditions. 4. Solubilized factor C, as well as membrane-bound factor C, could be alternatively stimulated by sodium taurocholate to give both glucocerebrosidase and ;acid' beta-glucosidase activities. 5. Membrane-bound factor C reacted optimally with factor P whereas solubilized factor C was preferentially stimulated by taurocholate. 6. Factor P-dependent glucocerebrosidase activity differed in kinetic properties from the taurocholate-stimulated enzyme activity. 7. The results are discussed in the light of (a) identity of glucocerebrosidase and ;acid' beta-glucosidase, (b) application in clinical diagnosis, (c) physiological significance of the enzyme system, and (d) polygenic inheritance in adult Gaucher's disease.  相似文献   

7.
Studies were done on the effect of bile salts on the rates of hydrolysis of the N-acetylneuraminyl linkages of several sialic acid-containing compounds by the neuraminidase of Clostridium perfringens. When GM3-ganglioside, two glycolipids (glycophorin and orosomucoid) and neuraminyl-lactose were used as substrates, hydrolysis was obtained even in the absence of bile salts, but addition of this detergent, below its critical micellar concentration, increased the reaction rates; above the critical micellar concentration of the detergent rates decreased again. When a second ganglioside, GM1, was used as substrate, the requirement for bile salts was absolute; hydrolysis was not observed at all without this detergent. With increasing concentrations of bile salt and in the presence of high concentrations of enzyme, rates of hydrolysis increased, reaching maximal values at fixed ratios of bile salt to GM1-ganglioside. Physical measurements showed that mixtures of bile salt and GM1-ganglioside form mixed micelles that have a higher critical micellar concentration, a lower molecular weight and greater axial ratio than the corresponding micelles of pure GM1-ganglioside.  相似文献   

8.
A major beta-glucosidase I and a minor beta-glucosidase II were purified from culture filtrates of the fungus Trichoderma reesei grown on wheat straw. The enzymes were purified using CM-Sepharose CL-6B cation-exchange and DEAE Bio-Gel A anion-exchange chromatography steps, followed by Sephadex G-75 gel filtration. The isolated enzymes were homogeneous in SDS-polyacrylamide gel electrophoresis and isoelectric focusing. beta-Glucosidase I (71 kDa) was isoelectric at pH 8.7 and contained 0.12% carbohydrate; beta-glucosidase II (114 kDa) was isoelectric at pH 4.8 and contained 9.0% carbohydrate. Both enzymes catalyzed the hydrolysis of cellobiose and p-nitrophenyl-beta-D-glucoside (pNPG). The Km and kcat/Km values for cellobiose were 2.10 mM, 2.45.10(4) s-1 M-1 (beta-glucosidase I) and 11.1 mM, 1.68.10(3) s-1 M-1 (beta-glucosidase II). With pNPG as substrate the Km and kcat/Km values were 182 microM, 7.93.10(5) s-1 M-1 (beta-glucosidase I) and 135 microM, 1.02.10(6) s-1 M-1 (beta-glucosidase II). The temperature optimum was 65-70 degrees C for beta-glucosidase I and 60 degrees C for beta-glucosidase II, the pH optimum was 4.6 and 4.0, respectively. Several inhibitors were tested for their action on both enzymes. beta-Glucosidase I and II were competitively inhibited by desoxynojirimycin, gluconolactone and glucose.  相似文献   

9.
Effects of metal ions on sphingomyelinase activity of Bacillus cereus   总被引:5,自引:0,他引:5  
Some divalent metal ions were examined for their effects on sphingomyelinase activity of Bacillus cereus. The enzyme activity toward mixed micelles of sphingomyelin and Triton X-100 proved to be stimulated by Co2+ and Mn2+, as well as by Mg2+. Km's for Co2+ and Mn2+ were 7.4 and 1.7 microM, respectively, being smaller than the Km for Mg2+ (38 microM). Sr2+ proved to be a competitive inhibitor against Mg2+, with a Ki value of 1 mM. Zn2+ completely abolished the enzyme activity at concentrations above 0.5 mM. The concentration of Zn2+ causing 50% inhibition of the enzyme activity was 2.5 microM. Inhibition by Zn2+ was not restored by increasing concentrations of Mg2+ when the concentration of Zn2+ was above 10 microM. Ba2+ was without effect. When sphingomyelinase was incubated with unsealed ghosts of bovine erythrocytes at 37 degrees C, the enzyme was significantly adsorbed onto the membrane in the presence of Mn2+, Co2+, Sr2+ or Ba2+. Incubation with intact or Pronase-treated erythrocytes caused enzyme adsorption only in the presence of Mn2+. In the course of incubation, the enzyme was first adsorbed on the membranes of intact bovine erythrocytes in the presence of Mn2+; then sphingomyelin breakdown proceeded with ensuing desorption of adsorbed enzyme. Hot-cold hemolysis occurred in parallel with sphingomyelin breakdown. In this case, the hydrolysis of membranous sphingomyelin as well as the initial enzyme adsorption took place in the following order: unsealed ghosts greater than Pronase-treated erythrocytes greater than intact erythrocytes.  相似文献   

10.
Thiophosphate analogs (C-S-P bond) of phosphatidylinositol (Cn-thio-PI: racemic hexadecyl-, dodecyl-, and octylthiophosphoryl-1-myo-inositol) and a fluorescent analog (pyrene-PI: rac-4-(1-pyreno)-butylphosphoryl-1-myo-inositol) were all substrates for phosphatidylinositol-specific phospholipase C from Bacillus cereus. Hydrolysis of thio-PI was followed by coupling the production of alkylthiol to a disulfide interchange reaction with dithiobispyridine. Hydrolysis of pyrene-PI was followed using a HPLC-based assay with fluorescence detection. The activity of PI-PLC with thio-PI analogs showed an interfacial effect. C16-Thio-PI, which had a critical micelle concentration (CMC) of 7 microM, gave a hyperbolic activity versus concentration curve between 0 and 2 mM, while C8-thio-PI, which had a CMC above 10 mM, showed very low activity which increased greatly upon introduction of an interface in mixed micelles with hexadecylphosphocholine (HDPC). Pyrene-PI, which aggregates above 0.3 mM, gave a sigmoidal activity curve with much higher activity above the CMC. All three thio-PI homologs as mixed micelles with HDPC gave hyperbolic activity curves with PI-PLC that were a function of bulk concentration of substrate at constant surface concentration and surface concentration of substrate at constant bulk concentration. The maximal activity of PI-PLC with pure C16-thio-PI micelles was 6.25 mumol min-1 mg-1, while that with pyrene-PI was estimated to be 68 mumol min-1 mg-1. With pure C16-thio-PI micelles, 0.022 mM substrate gave half Vmax, similar to that in mixed micelles with HDPC.  相似文献   

11.
AIMS: To compare endogenous and exogenous beta-glucosidases for the hydrolysis of the predominant isoflavone glucosides in soymilk in order to improve the biological activity. METHODS AND RESULTS: beta-glucosidase activity of probiotic organisms, including Bifidobacterium animalis ssp. lactis Bb12, Lactobacillus acidophilus ATCC 4461 and Lactobacillus casei 2607 in soymilk, was evaluated and was related to the increase in the concentration of isoflavone aglycones during fermentation. The concentrations of isoflavone compounds in soymilk were monitored using a Varian model HPLC with an Amperometric electrochemical detector. The aglycone composition, also known as aglycone equivalent ratio, has been considered to be important for the delivery of health benefits of isoflavones, and was monitored during the fermentation of soymilk. Comparison of the hydrolytic effectiveness of both exogenous and endogenous enzyme during 4-h incubation in soymilk was conducted using the Otieno-Shah (O-S) index. Results showed that exogenous enzyme exhibited faster rate of isoflavone glucoside hydrolysis than that by endogenous enzyme. Highest O-S indices were obtained after 4, 3 and 2 h of incubation with enzyme solution having beta-glucosidase activity of 0.288 U ml(-1), 0.359 U ml(-1) and 0.575 U ml(-1), resulting into aglycone concentration increments of 5.87-, 6.07- and 5.94-fold, respectively. Conversely, aglycone concentration in the soymilk with B. animalis ssp. lactis Bb12, L. casei 2607 and L. acidophilus 4461 increased by 3.43-, 2.72- and 3.03-fold, respectively, after 4 h of fermentation at 37 degrees C. In addition, the O-S index of endogenous enzyme was much lower than that of the exogenous enzyme over the same 4-h incubation period. Optimum aglycone equivalent ratios coincided with highest O-S indices and highest aglycone concentrations in soymilk hydrolysed with exogenous enzyme. The same correlation of O-S indices and highest aglycone concentrations occurred for endogenous enzyme during the 24 h of fermentation. CONCLUSIONS: Obtaining highest aglycone concentration and optimum aglycone equivalent ratio could provide a critical beginning point in clinical trials for the realization of unique health benefits of soy isoflavones. SIGNIFICANCE AND IMPACT OF THE STUDY: Screening for beta-glucosidase activities of probiotics in soymilk and comparing their hydrolytic potentials with that of exogenous beta-glucosidase could find wide applications in the development of different aglycone-rich functional soy beverages.  相似文献   

12.
The mechanism of herpesviral protease activation upon dimerization was studied using two independent spectroscopic assays augmented by directed mutagenesis. Spectroscopic changes, attributable to dimer interface conformational plasticity, were observed upon dimerization of Kaposi's sarcoma-associated herpesvirus protease (KSHV Pr). KSHV Pr's dissociation constant of 585 +/- 135 nM at 37 degrees C was measured by a concentration-dependent, 100-fold increase in specific activity to a value of 0.275 +/- 0.023 microM product min(-1) (microM enzyme)(-1). A 4 nm blue-shifted fluorescence emission spectrum and a 25% increase in ellipticity at 222 nm were detected by circular dichroism upon dimer association. This suggested enhanced hydrophobic packing within the dimer interface and/or core, as well as altered secondary structures. To better understand the structure-activity relationship between the monomer and the dimer, KSHV Pr molecules were engineered to remain monomeric via substitution of two separate residues within the dimer interface, L196 and M197. These mutants were proteolytically inactive while exhibiting the spectroscopic signature and thermal stability of wild type, dissociated monomers (T(M) = 75 degrees C). KSHV Pr conformational changes were found to be relevant in vivo, as the autoproteolytic inactivation of KSHV Pr at its dimer disruption site [Pray et al. (1999) J. Mol. Biol. 289, 197-203] was detected in viral particles from KSHV-infected cells. This characterization of structural plasticity suggests that the structure of the KSHV Pr monomer is stable and significantly different from its structure in the dimer. This structural uniqueness should be considered in the development of compounds targeting the dimer interface of KSHV Pr monomers.  相似文献   

13.
We have investigated effects of temperature on the catalytic and allosteric properties of the cGMP-stimulated cyclic nucleotide phosphodiesterase from calf liver. Vmax for cAMP and cGMP increased as assay temperature increased from 5 to 45 degrees C. At substrate concentrations below Kmapp, however, hydrolysis increased as temperature decreased from 45 to 5 degrees C and was much greater at 5 degrees C than at 45 degrees C. As assay temperature decreased, Kmapp for cAMP and cGMP decreased. Hill coefficients for cAMP and cGMP were approximately 1.9 at 45 degrees C and 1.2-1.0 at 5 degrees C. cGMP stimulated hydrolysis of 0.5 microM [3H]cAMP at all assay temperatures. Although maximal activity stimulated by cGMP, like Vmax, was lowest at 5 degrees C, presumably because of the effect of temperature on catalytic activity, the apparent activation constant (K alpha app) for cGMP stimulation was lower at 5 degrees C than at 45 degrees C. Thus, affinity for both substrate and effector was increased at 5 degrees C, suggesting that low temperature promotes transitions of the cGMP-stimulated phosphodiesterase to a "high affinity" state. That cGMP stimulated cAMP hydrolysis at 5 degrees C suggests that temperature-induced transitions are incomplete and/or readily reversible. In assays at 30 degrees C competitive inhibitors, like substrates, induce allosteric transitions which result in enhanced hydrolysis of low substrate (1.0 microM [3H] cAMP) concentrations. At higher substrate concentrations (50 microM [3H]cAMP), with the enzyme in the "activated" state, inhibitors compete with substrate at catalytic sites and reduce hydrolysis. At 45 degrees C, as at 30 degrees C, 1-methyl-3-isobutylxanthine (IBMX) and papaverine increased hydrolysis of 1.0 microM [3H]cAMP and reduced hydrolysis of 50 microM [3H]cAMP. At 5 degrees C, however, IBMX and papaverine inhibited hydrolysis of both 1.0 and 50 microM [3H]cAMP. Enzyme activity was relatively more sensitive to inhibition by IBMX at 5 degrees C than at 45 degrees C. Taken together, these observations support the notion that low temperature induces incomplete or readily reversible transitions to the high affinity state for substrates, effectors, and inhibitors. These observed effects of temperature also point out that enzyme determinants and topographical features responsible for transitions to the high affinity state and expression of catalytic activity can be regulated independently.  相似文献   

14.
The activation of docosahexaenoic acid by rat brain microsomes was studied using an assay method based on the extraction of unreacted [1-14C]docosahexaenoic acid and the insolubility of [1-14C]docosahexaenoyl-CoA in heptane. This reaction showed a requirement for ATP, CoA, and MgCl2 and exhibited optimal activity at pH 8.0 in the presence of dithiothreitol and when incubated at 45 degrees C. The apparent Km values for ATP (185 microM), CoA (4.88 microM), MgCl2 (555 microM) and [1-14C]docosahexaenoic acid (26 microM) were determined. The presence of bovine serum albumin or Triton X-100 in the incubation medium caused a significant decrease in the Km and Vm values for [1-14C]docosahexaenoic acid. The enzyme was labile at 45 degrees C (t1/2:3.3 min) and 37 degrees C (t1/2:26.5 min) and lost 36% of its activity after freezing and thawing. The transition temperature (Tc) obtained from Arrhenius plot was 27 degrees C with the activation energies of 74 kJ/mol between 0 degrees C and 27 degrees C and 30 kJ/mol between 27 degrees C and 45 degrees C. [1-14C]Palmitic acid activation in rat brain and liver microsomes showed apparent Km values of 25 microM and 29 microM respectively, with V values of 13 and 46 nmol X min-1 X mg protein-1. The presence of Triton X-100 (0.05%) in the incubation medium enhanced the V value of the liver enzyme fourfold without affecting the Km value. Brain palmitoyl-CoA synthetase, on the other hand, showed a decreased Km value in the presence of Triton X-100 with unchanged V. The Tc obtained were 25 degrees C and 28 degrees C for brain and liver enzyme with an apparent activation energy of 109 and 24 kJ/mol below and above Tc for brain enzyme and 86 and 3.3 kJ/mol for liver enzyme. The similar results obtained for the activation of docosahexaenoate and palmitate in brain microsomes suggest the possible existence of a single long-chain acyl-CoA synthetase. The differences observed in the activation of palmitate between brain and liver microsomes may be due to organ differences. Fatty acid competition studies showed a greater inhibition of labeled docosahexaenoic and palmitic acid activation in the presence of unlabeled unsaturated fatty acids. The Ki values for unlabeled docosahexaenoate and arachidonate were 38 microM and 19 microM respectively for the activation of [1-14C]docosahexaenoate. In contrast, the competition of unlabeled saturated fatty acids for activation of labeled docosahexaenoate is much less than that for activation of labeled palmitate.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

15.
Adenosine deaminase was purified (780-fold) from skeletal muscle of camel (Camelus Dormedarius) to homogeneity level by using DEAE Sephadex chromatography, ammonium sulfate precipitation, gel filtration and ion exchange chromatography. The enzyme appeared to be monomeric with subunit molecular weight of 43kDa and isoelectric point of 4.85. The enzyme showed specificity for adenosine and exhibited Michaelis-Menten Kinetics with kappa(cat) of 1112.41 min(-1) and K(m) of 14.7 microM at pH 7.5. The pH and temperature optima for enzyme activity were 7-7.5 and 25 degrees C, respectively. Free energy (DeltaG*), enthalpy (DeltaH*) and entropy (DeltaS*) of activation for denaturation of adenosine deaminase at 50 degrees C were 88.94, 99.65 kJmol(-1) and 33.16 Jmol(-1), respectively. The purified enzyme had half-lives of 636 and 61 min at 25 and 50 degrees C, respectively. The activation energy for catalysis of camel skeletal muscle adenosine deaminase was 9.13 kJmol(-1). Free energy (DeltaG#), enthalpy (DeltaH#) and entropy (DeltaS#) of activation for hydrolysis of adenosine deaminase at 25 degrees C were 50.35, 6.65 kJmol(-1) and -146.62 Jmol(-1), respectively. Purine riboside inhibited the enzyme competitively with K(i) of 16 microM.  相似文献   

16.
Spectrophotometric measurements of bilirubin-IX alpha in water and in aqueous/organic solvent mixtures at pH 10.0 as a function of bilirubin-IX alpha concentration (approx. 0.6--400 microM) are consistent with the formation of dimers (KD - 1.5 microM) in dilute (less than 10 microM) aqueous solution and further self-aggregation to multimers at higher concentrations. Added urea (to 10M) and increases in temperature (to 62 degrees C) obliterate the dimer-multimer transition at 10 microM, but added NaCl (to 0.30 M) promotes strong aggregation of dimers over a narrow concentration range, suggesting a 'micellization' phenomenon. Concentrations of dioxan or ethanol greater than 60% (v/v) in water were required to obtain the absorption spectrum of bilirubin-IX alpha monomers, suggesting that both hydrophobic and electrostatic (pi-orbital) interactions are involved in stabilizing the dimeric state in water. Micellar concentrations of sodium dodecyl sulphate induced spectrophotometric shifts in the dimer absorption spectrum of bilirubin-IX alpha consistent with progressive partitioning of bilirubin-IX alpha monomers into a relatively non-polar region of the micelles and allowed a deduction of the apparent critical micellar concentration that closely approximated the literature values. The pattern of bilirubin IX alpha association with bile salts is complex, since the absorption spectrum shifts hypsochromically below and bathochromically above the critical micellar concentration of the bile salts. Consistent with these observations, bilirubin IX alpha appears to bind to the polar face of bile salt monomers and to the polar perimeter of small bile salt micelles. At higher bile salt concentrations some-bilirubin-IX alpha monomers partition into the hydrophobic interior of the bile salt micelles. Our results suggest that under physiological conditions the natural conjugates of bilirubin-IX alpha may exhibit similar physical chemical properties in bile, in that dimers, highly aggregated multimers and bile salt-associated monomers may co-exist.  相似文献   

17.
A novel fluorescent phospholipid analogue, 1-triacontanoyl-2-(pyren-1-yl)hexanoyl-sn-glycero-3-phosphocholine (C30PHPC) was employed as a substrate for human pancreatic phospholipase A2. C30PHPC has a main endothermic phase transition with Tm at 46 degrees C as determined by differential scanning calorimetry (DSC). For an aqueous dispersion of C30PHPC the ratio of the intensities of pyrene excimer and monomer fluorescence emission, (IE/IM) has a maximum between 32 and 36 degrees C. The excimer emission intensity (at 480 nm) exceeds the monomer emission intensity (at 400 nm) 6.5-fold thus indicating a close packing of the phospholipid pyrene moieties in the lipid phase. C30PHPC has a limiting mean molecular area of 37 A2 at surface pressure 35 dyn cm-1 as judged by the compression isotherm at an air-water interphase. The hydrolysis of C30PHPC by human pancreatic phospholipase A2 was followed by monitoring the increase in the pyrene monomer fluorescence emission intensity occurring as a consequence of transfer of the reaction product, pyren-1-yl hexanoic acid into the aqueous phase. The enzyme reaction exhibited an apparent Km of 2.0 microM substrate. Calcium at a concentration of 0.2 mM activated the enzyme 4-fold. Maximal hydrolytic rates were obtained at 45 degrees C and at pH between 5.5 and 6.5. The enzyme reaction could be inhibited by 5 mM EDTA, confirming the absolute requirement for Ca2+ of this enzyme. The present fluorimetric assay easily detects hydrolysis of C30PHPC in the pmol min-1 range. Accordingly, less than nanogram levels of human pancreatic phospholipase A2 can be detected.  相似文献   

18.
Mixed acyl chain phosphatidylcholine molecules in Triton N-101 micelles were employed as substrates for lipoprotein lipase to test which substrate acyl chain has the greatest effect on activation of the enzyme by apolipoprotein C-II. The phospholipase A1 activity of lipoprotein lipase was measured by pH-stat. The activation factor (lipoprotein lipase activity plus apolipoprotein C-II/activity minus apolipoprotein C-II) increased monotonically with apolipoprotein C-II concentration up to 1 microM apolipoprotein C-II at an enzyme concentration of 0.01 microM. The maximal activation factor for phosphatidylcholine substrate molecules with sn-2 acyl chain lengths of 14 averages 14.8. By contrast, for sn-2 acyl chain lengths of 16 the activation factor was 29.2. Varying the sn-1 acyl chain length had no significant effect on the activation factor. The chain-length dependence of the activation factor is similar with the apolipoprotein C-II peptide fragment comprising residues 56-79, which does not include the lipid-binding region of apolipoprotein C-II. These data are consistent with a model for activation of lipoprotein lipase in which residues 56-79 bind to lipoprotein lipase and alter the interaction of the sn-2 acyl chain of the phosphatidylcholine (PC) substrate or the lysoPC product within the activated state complex.  相似文献   

19.
The ionization behavior of retinoic acid (RA) in an aqueous phase and when bound to bovine serum albumin was studied. Titrations of RA in the various phases were followed by monitoring the red shift in the absorption maximum of RA that occurred upon deprotonation. The apparent pK of RA was dependent on the concentration of this compound. At the concentration range 6-20 microM, the pK of RA in water had a value of approximately 8.0. As the concentration was decreased in the range 1-6 microM, the value of the pK decreased continuously. The lowest pK observed was approximately 6.0. It was concluded that RA in an aqueous phase at concentrations in the microM range, forms micelles, and that the values of the pK of RA monomers and micelles in water are less than 6.0 and 8.0, respectively. The presence of 0.15 M NaCl caused a decrease in the pK of RA micelles and lowered the value of the CMC. Titration of RA in the presence of bovine serum albumin revealed the presence of a heterogeneous population comprised of three distinct microenvironments for RA associated with this protein. Two populations of RA were found to undergo complete titration in the pH range 4-8. A third population became apparent at pH greater than 9.5.  相似文献   

20.
Lysophosphatidylcholine (lysoPC) monomers or micelles in water give rise to a narrow, isotropic phosphorus-31 NMR signal (40.6 ppm; v1/2 1.7 Hz; 32.2 MHz). Upon addition of praseodymium ions, the phosphorus signals are shifted downfield. However, the downfield shifts for the longer-chain lysophosphatidylcholines, which exist in the aggregated state, are far greater than those for the shorter-chain homologues, which exist as monomers. At a Pr3+/lysoPC molar ratio of 0.5, the signals of C12lysoPC through C18lysoPC were shifted by 12.1 ppm, whereas the signals of C6lysoPC and C8lysoPC were shifted by only 2.26 ppm. This very pronounced difference in lanthanide-induced downfield shifts between micelles and monomers can be utilized to determine with accuracy lysoPC critical micelle concentrations (CMC) from downfield shift-vs.-concentration plots. The CMC values we determined were 57 mM for C8lysoPC, 5.7 mM for C10lysoPC, and 0.6 mM for C12lysoPC. The shift reagent phosphorus-31 nuclear magnetic resonance technique particularly lends itself to the measurement of CMC values in the millimolar and high micromolar range. The method can equally be used for measuring critical micelle concentrations of short-chain phosphatidylcholines.  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