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 31 毫秒
1.
1. An acid aspartic proteinase in the regressing tadpole tail was purified about 800-fold with a 36% recovery. 2. The mol. wt of the enzyme was found to be 42,000 on gel filtration and 38,000 on sodium dodecyl sulfate polyacrylamide gel electrophoresis, respectively. 3. The purified enzyme had a maximum activity at pH 3.5 and an apparent Km of 0.084% with acid-denatured hemoglobin as substrate. 4. The enzyme activity was strongly inhibited by pepstatin. In addition, diazoacetylnorleucine methyl ester inactivated the enzyme in the presence of cupric ions. 5. The enzyme was identified as a cathepsin D (EC. 3.4.23.5)-like proteinase.  相似文献   

2.
Mucor pusillus acid protease was rapidly inactivated with 1 : 1 stoichiometry by reaction with diazoacetyl-DL-norleucine methyl ester (DAN) in the presence of cupric ions. Cupric ions were essential for this inactivation. The rate of inactivation was maximal at around pH 6 when the enzyme was mixed with DAN and cupric ions without prior mixing of the reagents, and at pH 5.3 when DAN and cupric ions were mixed and incubated before addition to the enzyme solution. In both cases, the rate of inactivation decreased as the pH was either increased or decreased. The amino acid composition of an acid hydrolysate of the DAN-Modified enzyme was indistinguishable from that of the native enzyme except for the incorporation of about one norleucine residue per molecule of protein. The enzyme was also inactivated by reaction with 1,2-epoxy-3-(p-nitrophenoxy)-propane (EPNP). At the stage of about 90% inactivation, 1.50 residues of EPNP were incorporated per molecule of protein and the rate of inactivation followed pseudo-first order kinetics. The optimal pH for the inactivation was pH 3.0 and the rate of inactivation decreased as the pH was either increased or decreased. Furthermore, the enzyme was strongly inhibited by pepstatin, and the reactions of DAN and of EPNP was also inhibited significantly by prior treatment of the enzyme with pepstatin. These results suggest that the enzyme may have two essential carboxyl groups at the active site, one reactive with DAN in the presence of cupric ions and the other with EPNP, and that pepstatin binds part of the active site to inhibit the reactions with DAN and EPNP as well as the enzyme activity.  相似文献   

3.
Comparative interactions of factor IX and factor IXa with human platelets   总被引:10,自引:0,他引:10  
Both factor IX and factor IXa were bound to gel filtered platelets in the presence of CaCl2 (2-20 mM) and human alpha-thrombin (0.06-0.2 units/ml) with maximal binding occurring in 10-20 min at 37 degrees C, and rapid reversibility was observed when unlabeled ligands were added in 100-fold molar excess. Competition studies with various coagulation proteins revealed that neither factor XI nor high molecular weight kininogen, at 300-fold molar excess, could compete with 125I-labeled factor IXa for binding sites on thrombin-activated platelets, whereas prothrombin and factor X, in 450-fold molar excess, could displace approximately 15 and 35%, respectively, of bound factor IXa in the absence of added factor VIII. Analysis of saturation binding data in the presence of CaCl2 and thrombin without factors VIII and X indicated the presence of 306 (+/- 57) binding sites per platelet for factor IX (Kd(app) = 2.68 +/- 0.25 nM) and 515 (+/- 39) sites per platelet for factor IXa (Kd = 2.57 +/- 0.14 nM). In the presence of thrombin-activated factor VIII (1-5 units/ml) and factor X (0.15-1.5 microM), the number of sites for factor IX was 316 (+/- 50) with Kd = 2.44 (+/- 0.30) nM and for factor IXa 551 (+/- 48) sites per platelet (Kd = 0.56 +/- 0.05 nM). Studies of competition for bound factor IXa by excess unlabeled factor IX or factor IXa, and direct 125I-labeled factor IXa binding studies in the presence of large molar excesses of factor IX, confirmed the conclusion from these studies that factor IX and factor IXa share approximately 300 low-affinity binding sites per thrombin-activated platelet in the presence of Ca2+ and in the absence of factor VIII and factor X, with an additional 200-250 sites for factor IXa with Kd(app) similar to that for factor IX. The presence of factor VIII and factor X increases by 5-fold the affinity of receptors on thrombin-activated platelets for factor IXa that participate in factor X activation.  相似文献   

4.
L D Faller 《Biochemistry》1990,29(13):3179-3186
TNP-ATP binds to the gastric H,K-ATPase with a 4.6-fold increase in fluorescence intensity and 10-nm blue shift that indicate a relatively hydrophobic protein environment. The fluorescence enhancement saturates and is compatible with binding to a single class of specific nucleotide sites with Kd less than 25 nM and N = 3.4 +/- 0.9 nmol mg-1. Cofactors of the H,K-ATPase affect the fluorescence enhancement. K+ causes a rapid fluorescence quench by binding to a single class of sites with Kd = 3 mM. Mg2+ rapidly and completely reverses the K+ quench and then causes a slow fluorescence quench. The maximum enhancement is approximately halved by either Mg2+ or K+ in titrations with both protein and fluorophore. Therefore, TNP-ATP reports changes in protein environment compatible with cofactor-induced changes in the conformation of the enzyme.  相似文献   

5.
N-(Benzyloxycarbonyl)-L-aspartyl-L-phenylalanine methyl ester (Z-AspPheOMe), a precursor of the synthetic sweetner asparatame, was synthesized from N-(benzyloxycarbolyl)-L-aspartic acid (Z-Asp) and L-phenylalanine methyl ester (PheOMe) with an immobilized thermolysin in various organic solvents. We found that in tert-amyl alcohol containing a small amount of water the immobilized enzyme showed a high activity comparble to that in ethyl acetate with quite a high stability. The immobilized enzyme was fully stable up to 70 degrees C in tert-amyl alcohol in the absence of the subatrate, and up to 50 degrees C in the presence of the substrate. The high stability in the presence of the substrate was found due to the fact that the release of calcium ions, the stabilizing factor of thermolysin, is suppressed.The substrate concentration dependence of the initial synthetic rate with the immobilized enzyme was quite different from that with the free enzyme in the biphasic system, in contrast to that in ethyl acetate. Finally, Z-AspPheOMe was continuously synthesized in a column reactor using 200 mM PheOMe and 120 mM Z-Asp as the substrate for over 300 h at 45 degrees C and a space velocity of 1 h(-1) without any loss of acivity. (c) 1994 John Wiley & Sons, Inc.  相似文献   

6.
1. The Type B acid protease from Aspergillus niger var. macrosporus was inactivated by reaction with diazoacetyl-DL-norleucine methyl ester (DAN), DL-1-diazo-3-tosylamido-2-heptanone (DTH), and L-1-diazo-3-tosylamido-4-phenyl-2-butanone (DTPB) in the presence of cupric ions. The reaction with DAN took place with 1:1 stoichiometry. The enzyme was also inactivated by reaction with 1, 2-epoxy-3-(p-nitrophenoxy)-propane (EPNP) with concomitant incorporation of approximately two EPNP molecules per molecule of protein. Moreover, these reactions of DAN and of EPNP were markedly inhibited by pepstatin. These results seem to indicate that, as in the case of porcine pepsin [EC 3.4.23.1] and related acid proteases, the enzyme has two essential carboxyl groups at the active site, one reactive with DAN and related diazo reagents in the presence of cupric ions and the other reactive with EPNP, and that pepstatin binds in the vicinity of these residues. 2. The Type A acid protease from the same mold, on the other hand, was found to be markedly less sensitive to these specific inhibitors. Under conditions where the Type B enzyme was completely inactivated by DAN and related diazo reagents, only partial inactivation of this enzyme occurred. The effect of prior mixing of DAN and cupric ions on the pH profile of inactivation was also different from that for the Type B enzyme. Moreover, the Type A enzyme was not inactivated by EPNP. These results thus indicate that the nature of the active site of the Type A enzyme is rather different from that of the Type B enzyme and hence that the Type A enzyme belongs to a different class of acid proteases from the Type B enzyme.  相似文献   

7.
The binding of glucono-1,5-lactone (gluconolactone) with almond beta-glucosidase was studied at pH 5.0 and 25 degrees C, in the absence and presence of glucose, by monitoring the enzyme fluorescence as a probe. From the results of fluorometric titration, the dissociation constant Kd and the maximum fluorescence intensity increase (percent) of the enzyme-gluconolactone complex relative to the enzyme alone, delta Fmax, were determined to be 12.7 microM and 14.7%, respectively. From the study of the temperature dependence of Kd, delta G degrees, delta H degrees and delta S degrees for the binding were evaluated to be -6.7 kcal mol-1, -3.5 kcal mol-1, and 10.8 e.u. (cal mol-1 deg-1), respectively, at 25 degrees C. The analysis of the fluorometric titration data in the presence of glucose revealed that these ligands bind competitively to the enzyme, probably at the same site. The results of a stopped-flow kinetic study are consistent with the following two-step mechanism: (formula; see text) which indicates that gluconolactone (L) and the enzyme (E) transiently form a loosely bound complex, ELtr (k-1/k+1 = 4.5 mM), in the first rapid bimolecular association step, and ELtr is converted into a more tightly bound complex EL (k+2 = 94 s-1, k-2 = 0.36 s-1) in the subsequent slow unimolecular process. The fluorescence intensity increase occurs solely in the latter step.  相似文献   

8.
Skene IK  Brooker JD 《Anaerobe》1995,1(6):321-327
A strain of the anaerobe Selenomonas ruminantium subsp. ruminantium that is capable of growing on tannic acid or condensed tannin as a sole energy source has been isolated from ruminal contents of feral goats browsing tannin-rich Acacia sp. Growth on tannic acid was accompanied by release of gallic acid into the culture medium but the bacterium was incapable of using gallic acid as a sole energy source. Tannin acylhydrolase (EC 3.1.1.20) activity was measured in crude cell-free extracts of the bacterium. The enzyme has a pH optimum of 7, a temperature optimum of 30-40 degrees C and a molecular size of 59 kDa. In crude extracts, the maximal rate of gallic acid methyl ester hydrolysis was 6.3 micromol min(-1) mg(-1) of protein and the K(m) for gallic acid methyl ester was 1.6 mM. Enzyme activity was displayed in situ in polyacrylamide and isoelectric focusing gels and was demonstrated to increase 17-fold and 36-fold respectively when cells were grown in the presence of gallic acid methyl ester or tannic acid.  相似文献   

9.
Crystalline L-histidine ammonia-lyase of Achromobacter liquidum was prepared with a 24% recovery of the activity. The specific activity of the pure enzyme (63 mumol of urocanic acid min-1 mg-1) is similar to those so far reported for the enzyme from other sources. The purified enzyme appeared to be homogeneous by analytical disc electrophoresis and isoelectric focusing (pI = 4.95). The molecular weight determined by Sephadex G-200 gel filtration is 200000. The optimum pH is 8.2, and the optimum temperature is 50 degrees C. The enzyme showed strict specificity to L-histidine (Km = 3.6 mM). Several histidine derivatives are not susceptible to the enzyme but do inhibit the enzyme activity competitively; the most effective inhibitors are L-histidine methyl ester (Ki = 3.66 mM) and beta-imidazole lactic acid (Ki = 3.84 mM). L-Histidine hydrazide (Ki = 36 mM) and imidazole (Ki = 6 mM) noncompetitively inhibited the enzyme EDTA markedly inhibited enzyme activity and this inhibition were reversed by divalent metal ions such as Mn2+, Co2+ Zn2+, Ni2+, Mg2+, and Ca2+. These results suggest that the presence of divalent metal ions is necessary for the catalytic activity of histidine ammonia-lyase. Sodium borohydride and hydrogen peroxide inhibited the enzyme activity.  相似文献   

10.
A M Hanel  W P Jencks 《Biochemistry》1991,30(47):11320-11330
The internalization of 45Ca by the calcium-transporting ATPase into sarcoplasmic reticulum vesicles from rabbit muscle was measured during a single turnover of the enzyme by using a quench of 7 mM ADP and EGTA (25 degrees C, 5 mM MgCl2, 100 mM KCl, 40 mM MOPS.Tris, pH 7.0). Intact vesicles containing either 10-20 microM or 20 mM Ca2+ were preincubated with 45Ca for approximately 20 s and then mixed with 0.20-0.25 mM ATP and excess EGTA to give 70% phosphorylation of Etot with the rate constant k = 300 s-1. The two 45Ca ions bound to the phosphoenzyme (EP) become insensitive to the quench with ADP as they are internalized in a first-order reaction with a rate constant of k = approximately 30 s-1. The first and second Ca2+ ions that bind to the free enzyme were selectively labeled by mixing the enzyme and 45Ca with excess 40Ca, or by mixing the enzyme and 40Ca with 45Ca, for 50 ms prior to the addition of ATP and EGTA. The internalization of each ion into loaded or empty vesicles follows first-order kinetics with k = approximately 30 s-1; there is no indication of biphasic kinetics or an induction period for the internalization of either Ca2+ ion. The presence of 20 mM Ca2+ inside the vesicles has no effect on the kinetics or the extent of internalization of either or both of the individual ions. The Ca2+ ions bound to the phosphoenzyme are kinetically equivalent. A first-order reaction for the internalization of the individual Ca2+ ions is consistent with a rate-limiting conformational change of the phosphoenzyme with kc = 30 s-1, followed by rapid dissociation of the Ca2+ ions from separate independent binding sites on E approximately P.Ca2; lumenal calcium does not inhibit the dissociation of calcium from these sites. Alternatively, the Ca2+ ions may dissociate sequentially from E approximately P.Ca2 following a rate-limiting conformational change. However, the order of dissociation of the individual ions can not be distinguished. An ordered-sequential mechanism for dissociation requires that the ions dissociate much faster (k greater than or equal to 10(5) s-1) than the forward and reverse reactions for the conformational change (k-c = approximately 3000 s-1). Finally, the Ca2+ ions may exchange their positions rapidly on the phosphoenzyme (kmix greater than or equal to 10(5) s-1) before dissociating. A Hill slope of nH = 1.0-1.2, with K0.5 = 0.8-0.9 mM, for the inhibition of turnover by binding of Ca2+ to the low-affinity transport sites of the phosphoenzyme was obtained from rate measurements at six different concentrations of Mg2+.  相似文献   

11.
Extracellular alpha-galactosidase, a glycoprotein from the extracellular culture fluid of Aspergillus ficuum grown on glucose and raffinose in a batch culture system, was purified to homogeneity in five steps by ion exchange and hydrophobic interaction chromatography. The molecular mass of the enzyme was 70.8 Kd by SDS polyacrylamide gel electrophoresis and 74.1 Kd by gel permeation HPLC. On the basis of a molecular mass of 70.7 Kd, the molar extinction coefficient of the enzyme at 279 nm was estimated to be 6.1 X10(4) M-1 cm-1. The purified enzyme was remarkably stable at 0 degrees C. It had a broad temperature optimum and maximum catalytic activity was at 60 degrees C. It retained 33% of its activity after 10 min. at 65 degrees C. It had a pH optimum of 6.0. It retained 62% of its activity after 12 hours at pH 2.3. The Kms for p-nitrophenyl-alpha-D-galactopyranoside, o-nitrophenyl-alpha-D-galactopyranoside and m-nitrophenyl-alpha-D-galactopyranoside are: 1462, 839 and 718 microM. The enzyme was competitively inhibited by mercury (19.8 microM), silver (21.5 microM), copper (0.48 mM), zinc (0.11 mM), galactose (64.0 mM) and fructose (60.3 mM). It was inhibited non-competitively by glucose (83.2 mM) and uncompetitively by mannose (6.7 mM).  相似文献   

12.
The affinity of phosvitin with serine hydroxymethyl transferase (SHMT), an acidic multi-subunit protein, was evaluated by measurements of enzyme activity, sedimentation velocity, steady-state fluorescence, circular dichroism and kinetic thermal stability. While the presence of phosvitin had no effect on the SHMT activity, the sedimentation coefficient of SHMT increased from 8.7 S to 12.5 S suggesting the formation of a complex at a SHMT:phosvitin molar ratio of 2:1. Based on steady-state fluorescence quenching measurements an association constant of 2.4 +/- 0.2 x 10(5) M-1 at 25 degrees C was obtained for the interaction of phosvitin with SHMT. The temperature dependency of the association constant in the range 15-35 degrees C suggests the involvement of ionic forces in the interaction. The thermal inactivation of SHMT followed first order kinetics. In the presence of phosvitin the rate constant decreased and half time increased. The circular dichroism measurements suggest that phosvitin interaction does not involve pyridoxal phosphate binding domain of the enzyme. Although minor changes in the secondary structure of the enzyme were observed, the environment around aromatic amino acids did not change significantly.  相似文献   

13.
cGMP-dependent protein kinase binds 4 mol cGMP/mol enzyme to two different sites. Binding to site 1 (apparent Kd 17 nM) shows positive cooperativity and is inhibited by Mg . ATP, whereas binding to site 2 (apparent Kd 100-150 nM) is non-cooperative and not affected by Mg . ATP. Autophosphorylation of the enzyme abolishes the cooperative binding to site 1 and the inhibitory effect of Mg . ATP. The association (K1) and dissociation (K-1) rate constant for site 2 and K1 for site 1 are not affected significantly by Mg . ATP or autophosphorylation. The dissociation rate from site 1 measured in the presence of 1 mM unlabelled cGMP is decreased threefold and over tenfold by Mg . ATP and autophosphorylation, respectively. In contrast, the dissociation rate from site 1 measured after a 500-fold dilution of the enzyme-ligand complex is 100-fold faster than that determined in the presence of 1 mM cGMP and is only slightly influenced by Mg . ATP or autophosphorylation. Only Kd values calculated with the latter K-1 values are similar to the Kd values obtained by equilibrium binding. These results suggest that autophosphorylation of cGMP-dependent protein kinase affects mainly the binding characteristics of site 1.  相似文献   

14.
The tetrahydropteroylglutamate methyltransferase from green beans (Phaseolus vulgaris) has been purified 80-fold by ion exchange chromatography and gel filtration. Optimal methyl transfer is found at pH 6.5 and 39 degrees C. Even at 0 degrees C, however, a considerable catalytic rate is observed. The Michaelis-Menten constants for homocysteine and 5-methyltetrahydropteroylglutamate are 0.43mM and 2.4 mM, respectively. Magnesium ions enhance the activity. Even purified preparations appear to contain traces of magnesium ions firmly bound, since a residual activity is found without addition of magnesium salts. Though the reaction requires anaerobiosis, an excess of reducing agents is inhibitory. The molecular weight of the transferase, determined by gel filtration, is 40 000 +/- 6%.  相似文献   

15.
Nucleotide binding to sarcoplasmic reticulum vesicles was investigated in the absence of calcium using both filtration and fluorescence measurements. Filtration assays of binding of radioactive nucleotides at concentrations up to 0.1 mM gave a stoichiometry of one ATP-binding site/sarcoplasmic reticulum ATPase molecule. When measured in the presence of calcium under otherwise similar conditions, ATPase velocity rose 4-8-fold (depending on pH and magnesium concentration) when the ATP concentration was increased from 1 microM to 0.1 mM. Binding of ATP and ADP enhanced the intrinsic fluorescence of sarcoplasmic reticulum ATPase, but AMP and adenosine did not affect it. Both filtration and fluorescence measurements showed that binding of metal-free ATP is independent of pH (Kd = 20-25 microM) but that the presence of magnesium induces pH dependence of the binding of the Mg.ATP complex (Kd = 10 microM at pH 6.0 and 1.5 microM at pH 8.0). Binding of metal-free ADP was pH-dependent but was not affected by magnesium. High magnesium concentrations inhibited nucleotide binding. These results suggest that ATP interacts with two different domains of Ca-ATPase that form the catalytic site. The first domain may bind the adenine moiety of the substrate, and the pH dependence of ADP binding suggests the participation of His683 in this region. The second domain of the catalytic site may bind the gamma-phosphate and the magnesium ion of the Mg.ATP complex and constitute the locus of the electrostatic interactions between the substrate and the enzyme.  相似文献   

16.
When 0.5 M sodium thiocyanate is added to uterine cytosol previously labeled with excess [3H]-17 beta-estradiol, no change can be detected in the steady-state cytosol concentration of [3H]estradiol-receptor complex for at least 20 h at 4 degrees C. However, the rate of exchange of bound estradiol in the presence of NaSCN was found to be substantially higher than that in the absence of the chaotropic salt. In the presence of NaSCN, the dissociation rate of the complex increases about 10-fold (K-1 SCN = 1.10 x 10(-2) min-1 vs. K-1 = 1.07 X 10 (-3)min-1) while the rate of association increases about 2-fold (K1 SCN = 1.2 X 10(7) min-1M-1 vs.K1= 7.4 X 10(6) min-1 M-1). The Kd changes 6.4-fold (Kd SCN = 9 X 10(-10) M vs. Kd = 1.4 x 10(-10 M) with no decrease in the number of binding sites as shown by Scatchard plots of saturation experiments. This effect of NaSCN can be exploited to assay preformed estrogen-receptor complex by exchange with [3H]estradiol at low temperature. When the sample containing preformed complex is incubated overnight (16 h) at 4 degrees C with excess [3H]estradiol in the presence of 0.5 M NaSCN, there is a quantitative exchange of nonlabeled for estradiol without loss of binding sites. Hormonal steroids other than estrogens do not interfere, and the exchange estradiol is bound with high affinity. Precision, accuracy, and linearity of the method are highly satisfactory.  相似文献   

17.
The interaction of vanadate ions with the Ca-ATPase from sarcoplasmic reticulum vesicles was studied in a native and a fluorescein-labeled ATPase preparation (Pick, U., and Karlish, S. J. D. (1980) Biochim. Biophys. Acta 626, 255-261). Vanadate induced a fluorescence enhancement in a fluorescein-labeled enzyme, indicating that it shifts the equilibrium between the two conformational states of the enzyme by forming a stable E2-Mg-vanadate complex (E2 is the low affinity Ca2+ binding conformational state of the sarcoplasmic reticulum Ca-ATPase). Indications for tight binding of vanadate to the enzyme (K1/2 = 10 microM) in the absence of Ca2+ and for a slow dissociation of vanadate from the enzyme in the presence of Ca2+ are presented. The enzyme-vanadate complex was identified by the appearance of a time lag in the onset of Ca2+ uptake and by a slowing of the fluorescence quenching response to Ca2+. Ca2+ prevented the binding of vanadate to the enzyme. Pyrophosphate (Kd = 2 mM) and ATP (Kd = 25 microM) competitively inhibited the binding of vanadate, indicating that vanadate binds to the low affinity ATP binding site. Binding of vanadate inhibited the high affinity Ca2+ binding to the enzyme at 4 degrees C. Vanadate also inhibited the phosphorylation reaction by inorganic phosphate (Ki = 10 microM) but had no effect on the phosphorylation by ATP. It is suggested that vanadate binds to a special region in the low affinity ATP binding site which is exposed only in the E2 conformation of the enzyme in the absence of Ca2+ and which controls the rate of the conformation transition in the dephosphorylated enzyme. The implications of these results to the role of the low affinity ATP binding sites are discussed.  相似文献   

18.
M T Mas  R F Colman 《Biochemistry》1985,24(7):1634-1646
Spectroscopic, ultrafiltration, and kinetic studies have been used to characterize interactions of reduced and oxidized triphosphopyridine nucleotides (TPNH and TPN), 2'-phosphoadenosine 5'-diphosphoribose (Rib-P2-Ado-P), and adenosine 2',5'-bisphosphate [Ado(2',5')P2] with with TPN-specific isocitrate dehydrogenase. Close similarity of the UV difference spectra and of the protein fluorescence changes accompanying the formation of the binary complexes provides evidence for the binding of these nucleotides to the same site on the enzyme. From the pH dependence of the dissociation constants for TPNH binding to TPN-specific isocitrate dehydrogenase in the absence and in the presence of Mn2+, over the pH range 5.8-7.6, it has been demonstrated that the nucleotide binds to the enzyme in its unprotonated, metal-free form. The involvement of positively charged residues, protonated over the pH range studied, has been postulated. One TPNH binding site per enzyme subunit has been measured by fluorescence and difference absorption titrations. A dramatic effect of ionic strength on binding has been demonstrated: about a 1000-fold decrease in the dissociation constant for TPNH has been observed at pH 7.6 upon decreasing ionic strength from 0.336 (Kd = 1.2 +/- 0.2 microM) to 0.036 M (Kd = 0.4 +/- 0.1 nM) in the presence and in the absence of 100 mM Na2SO4, respectively. Weak competition of sulfate ions for the nucleotide binding site has been observed (KI = 57 +/- 3 mM). The binding of TPN in the presence of 100 mM Na2SO4 at pH 7.6 is about 100-fold weaker (Kd = 110 +/- 22 microM) than the binding of the reduced coenzyme and is similarly affected by ionic strength. These results demonstrate the importance of electrostatic interactions in the binding of the coenzyme to TPN-specific isocitrate dehydrogenase. The large enhancement of protein fluorescence caused by binding of TPN and Rib-P2-Ado-P (delta Fmax = 50%) and of Ado(2',5')P2 (delta Fmax = 41%) has been ascribed to a local conformational change of the enzyme. An apparent stoichiometry of 0.5 nucleotide binding site per peptide chain was determined for TPN, Rib-P2-Ado-P, and Ado(2',5')P2 from fluorescence titrations, in contrast to one binding site per enzyme subunit determined from UV difference spectral titration and ultrafiltration experiments. Thus, the binding of one molecule of the nucleotide per dimeric enzyme molecule is responsible for the total increase in protein fluorescence, while binding to the second subunit does not cause further change.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Exo-(1,4)-alpha glucan lyase (GLase) was purified from a red alga Gracilaria chorda. The enzyme was activated 1.3-fold in the presence of Ca(2+) and Cl(-) ions. The ions also stabilized the enzyme increasing the temperature of its maximum activity from 45 degrees C to 50 degrees C. GLase was inactivated by chemical modification with carbodiimide and a carboxyl group of the enzyme was shown essential to the lyase activity. A tryptophanyl residue(s) was also shown to be important for the activity and was probably involved in substrate binding. K(m) values of the enzyme were 2.3 mM for maltose, 0.4 mM for maltotriose and 0.1 mM for maltooligosaccharides of degree of polymerization (dp) 4-7, and the k(0) values for the oligosaccharides were similar (42-53 s(-1)). The analysis of these kinetic parameters showed that the enzyme has four subsites to accommodate oligosaccharides. The subsite map of GLase was unique, since subsite 1 and subsite 2 have large positive and small negative affinities, respectively. The subsite map of this type has not been found in other enzymes with exo-action on alpha-1,4-glucan. The K(m) and k(0) values for the polysaccharides were lower (0.03 mM) and higher (60-100 s(-1)), respectively, suggesting the presence of another affinity site specific to the polysaccharides.  相似文献   

20.
This study systematically examined the characteristics of specific binding of adult diferric transferrin to its receptor using a Triton X-100 solubilized preparation from human placentas as the receptor source. The following information was obtained. The ionic strength for maximal binding is in the range of 0.1-0.3 M NaCl. The pH optimum for specific binding extends over the range, from pH 6.0-10.0. Specific binding of diferric transferrin is not affected by 2.5 approximately 50 mM CaCl2 or by 10 mM EDTA. Triton X-100 in the concentration range of 0.02-3.0% does not affect specific binding. Specific binding is saturated within 10 min at 25 or 37 degrees C in the presence of excess amounts of diferric transferrin. The binding is reversible and the dissociation of diferric transferrin from the transferrin receptor is complete within 40 min at 25 degrees C. Apotransferrin, both adult and fetal, showed less binding than the holotransferrin species by competitive binding assay in the presence of 10 mM EDTA independent of up to 20 mM CaCl2. A 1500-fold molar excess of adult and fetal apotransferrin is required to give 40% inhibition for 125I-labeled diferric transferrin binding. Since calcium ion is not a factor, and since apotransferrin has such high binding affinity for iron (Ka = 1 X 10(24], this experiment suggests that the EDTA was necessary to prevent conversion of apotransferrin to holotransferrin from available iron in the reaction system. The specificity of the transferrin receptor for transferrin was examined by competitive binding studies in which 125I-diferric transferrin binding was measured in the presence of a series of other proteins. The proteins tested in the competitive binding studies were classified into three groups; in the first group were human serum albumin and ovalbumin; in the second group were proteins containing iron ions, such as hemoglobin, hemoglobin-haptoglobin complex, heme-hemopexin complex, ferritin, and diferric lactoferrin; in the third group were the metal-binding serum proteins, ceruloplasmin and metallothionein. None of these proteins except ferritin showed inhibition of diferric transferrin binding to the receptor. The effect of ferritin was small since a 700- to 1500-fold molar excess of ferritin is required for 50% inhibition of binding of diferric transferrin to the receptor.  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