首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A novel extracellular serine protease designated Pernisine was purified to homogeneity and characterized from the archaeon Aeropyrum pernix K1. The molecular mass, estimated by SDS-PAGE analysis and by gel filtration chromatography, was about 34 kDa suggesting that the enzyme is monomeric. Pernisine was active in a broad range of pH (5.0-12.0) and temperature (60-120 degrees C) with maximal activity at 90 degrees C and between pH 8.0 and 9.0. In the presence of 1 mM CaCl(2) the activity, as a function of the temperature, reached a maximum at 90 degrees C but at 120 degrees C the enzyme retained almost 80% of its maximal activity. Activity inhibition studies suggest that the enzyme is a serine metalloprotease and biochemical data indicate that Pernisine is a subtilisin-like enzyme. The protease gene, identified from the sequenced genome of A. pernix, was amplified from total genomic DNA by PCR technique to construct the expression plasmid pGEX-Pernisine. The Pernisine, lacking the leader sequence, was expressed in Escherichia coli BL21 strain as a fusion protein with glutathione- S-transferase. The biochemical properties of the recombinant enzyme were found to be similar to those of the native enzyme.  相似文献   

2.
A bacterium that secreted cyclodextrin glucanotransferase (CGTase) in a medium overlaid with n-hexane was isolated and identified as Paenibacillus illinoisensis strain ST-12 K. The CGTase of the strain was purified from the culture supernatant. The molecular mass was 70 kDa. The enzyme was stable at pH 6 to 10 and active at pH 5.0 to 8.0. The optimum temperature at pH 7.0 was 65 degrees C in the presence of 5 mM CaCl2. The enzyme produced mainly beta-cyclodextrin. The total yield of alpha-, beta-, and gamma- cyclodextrins was increased 1.4-fold by the addition of ethanol. In particular, the yield of beta-cyclodextrins in the presence of 10% (vol/vol) ethanol was 1.6-fold that without ethanol. The CGTase was stable and active in the presence of large amounts of various organic solvents.  相似文献   

3.
Purification and characterization of the sea urchin embryo hatching enzyme   总被引:11,自引:0,他引:11  
The sea urchin hatching enzyme provides an interesting model for the control of gene expression during early development. In order to study its properties and developmental regulation, the hatching enzyme of the species Paracentrotus lividus has been purified. The fertilization envelopes of the embryos were digested before hatching by a crude culture supernatant previously made. The enzyme was then solubilized by 1 M NaCl and 0.5% 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate and purified by hydrophobic chromatography on Procion-agarose. A 470-fold increase in specific activity was obtained. The kinetic parameters of the proteolytic activity using dimethylcasein as substrate are: Km = 120 micrograms x ml-1, Vm = 200 mumol x min-1 x mg-1, and kcat = 180 s-1 at 500 mM NaCl, 10 mM CaCl2, pH 8.0, at 35 degrees C. The purified enzyme is highly active on fertilization envelopes: at 20 degrees C and 500 mM NaCl, 10 mM CaCl2, pH 8.0, 100 ng of enzyme completely denudes embryos in about 20 min under standard conditions. The molecular mass of the enzyme was estimated as 57 kDa by gel filtration, 51 kDa by gel electrophoresis, and 52 kDa by amino acid analysis. The hatching enzyme was shown to be a glycoprotein which autolyzes to a 30-kDa inactive form. Antibodies raised against the 51- or 30-kDa forms reacted with both these forms. Immunoblotting experiments showed that the hatching supernatants contain important amounts of the autolyzed species.  相似文献   

4.
Glyceraldehyde 3-phosphate (Glyc3P), a glycolytic intermediate, non-enzymatically glycosylated (or glycated) and inhibited the pig heart cytoplasmic aspartate aminotransferase (cAAT). Glyc3P (5.0 mM) decreased cAAT activity by 47% after 1 min at 23 degrees C. cAAT activity remained unchanged after a 24h incubation with either glucose 6-phosphate (5.0 mM) or ribose 5-phosphate (5.0 mM). Increasing the incubation pH from 6.4 to 7.8 or the incubation temperature from 23 degrees C to 50 degrees C enhanced Glyc3P's inhibitory effect on cAAT activity. Glyc3P (250-500 microM) decreased the thermal stability of cAAT as evidenced by lowering the Tm or temperature that caused a 50% irreversible loss of cAAT activity (69 degrees C, control; 58.5 degrees C, 500 microM Glyc3P). Glyc3P decreased cAAT amino group content and increased glycation products, which were measured by adduct formation, fluorescence and protein crosslinking.  相似文献   

5.
Recently, protease 2A of human rhinovirus 2 (HRV2 2A) was shown to require a zinc ion for the formation of an active enzyme although zinc is not involved mechanistically. The data presented clearly show that the zinc ion bound to a picornaviral-specific motif represents an essential component of the native structure, probably representing a new Zn-binding motif. This structure, containing mostly beta-strand elements as shown by CD spectroscopy, changes drastically upon removal of zinc. The zinc-depleted form does represent an intermediate with mostly unchanged secondary structure, but not a fully denatured random coil as obtained by guanidinium hydrochloride. This is indicated by the blue-shifted fluorescence spectra and by CD. The native protein exhibited a cooperative phase transition at 53 degrees C. In contrast, the zinc-depleted form did not show any transition at all, again demonstrating the stabilizing role of the zinc ion. A structural intermediate was observed during thermal and pH denaturation that may represent a molten globule, as suggested by its ANS binding.  相似文献   

6.
The secondary and tertiary structure of the oligomeric arginase (EC 3.5.3.1) from beef liver was investigated by circular dichroism (CD) and fluorescence measurements. The far-ultraviolet CD spectrum of the enzyme at neutral pH is indicative of high helical content. The intrinsic fluorescence emission of the protein is due to tryptophan, the contribution of tyrosine being small. Upon excitation at 295 nm, the maximum of emission occurs at 330 nm, implying that the tryptophan residues are rather buried in a hydrophobic interior of the protein. Ethylenediaminetetraacetic acid (EDTA), which inactivates the enzyme by removing the functional Mn2+-ion from the enzyme, does not dissociate the enzyme into subunits, nor affect noticeably its secondary and tertiary structure. Inactivation occurs in the acid pH range, being complete at pH below 4. However, acidification up to pH 1.5 produced only limited changes in the far-ultra-violet CD spectrum and intrinsic fluorescence emission properties. The enzyme shows noteworthy thermal stability, as shown by measuring the residual activity after heating and by evaluating the temperature dependence of the CD signal at 220 nm and the intensity of emission fluorescence. A temperature of half inactivation (Tm) of 77 degrees was determined upon heating the enzyme at pH 7.5 in the presence of Mn2+-ions for 10 min; in the presence of EDTA, Tm is shifted to 55 degrees. Taken together, these observations indicate that the structural stability of beef liver arginase arises from a clustering of hydrophobic amino acids and from Mn2+-ion binding.  相似文献   

7.
Hemolymph of decapod crustaceans contains lectins of important specificity. An isolation procedure, based on adsorption of hemolymph lectins on red blood cells (RBC) fixed with formaldehyde, is described. Hemolymph is let to clot for 3 h at 22-28 degrees C (RT) and for 24 h at 5 degrees C; centrifuged at 13000 g for 30 min; filtered through 5-microm filters; diluted with an equal volume of 50 mM NaCl, 100 mM CaCl(2); supplemented with protease as well as phenoloxidase inhibitors; centrifuged at 13000 g for 20 min. Formalinized RBC (FRBC) are mixed with diluted hemolymph to a suspension of about 20% v/v FRBC. After incubation for 30 min at RT, FRBC are washed five times with 150 mM NaCl, 10 mM CaCl(2). The lectins adsorbed on FRBC are desorbed using either 100-500 mM of carbohydrate solutions in 0.9% NaCl or 50 mM Tris-HCl buffer, pH 8.0 containing 100 mM NaCl and 20 mM entylenediaminetetraacetate (EDTA). The procedure is efficient in isolating the hemolymph lectins of the decapods Liocarcinus depurator and Potamon potamios.  相似文献   

8.
Effects of pH, enzyme concentration, and various supplements on the catalytic activity, temperature stability, and secondary structure of horseradish peroxidase (HRP) were studied in diluted aqueous solutions. In 5.0 mM citrate-phosphate buffer (pH 4.2) at 55 degrees C and infinite dilution, HRP was inactivated with a rate constant of 2.86 x 10(-3) s-1. CaCl2, BSA, and glycerol caused protective effects, whereas KCl, LiCl, maltose, PEG-6000 (at a concentration above 3%), Triton X-100, ethanol, and Kathon CG had an opposite effect and altered the secondary structure of HRP. Two HRP-stabilizing media: the "glycerol-based" one containing 10% ethanol and 20% glycerol, or the "protein-based" one containing 0.1% Kathon CG and 0.2 g/l of BSA in 50.0 mM Tris-HCl buffer (pH 7.2) supplemented with 50 mM CaCl2 were developed, and the stability of HRP (0.36 nM) and its immunoglobulin, cortisol, and progesterone conjugates were compared in these two media. The protein-based medium displayed a greater stabilizing effect particularly on HRP-steroid conjugates.  相似文献   

9.
R W Olafson  L B Smillie 《Biochemistry》1975,14(6):1161-1167
Streptomyces griseus trypsin has been isolated from Pronase by ion-exchange chromatography on CM-Sephadex and SE-Sephadex. The isolated enzyme was homogeneous by the criteria tested except for a low degree of contamination by an enzyme with nontryptic activity. The latter could be partially resolved by chromatography on Bio-Rex 70. The molar absorbancy at 280 nm was found to be 3.96 times 10-4 M-1/cm and the E1cm1% was found to be 17.3. The molecular weight was 22,800 plus or minus 800. The enzyme was found to be stable at 0 degrees from pH 2 to 10. At 30 degrees the enzyme was maximally stable at pH 3-4 and significantly stabilized in the neutral and alkaline range by 15 mM Ca2+. Some evidence was obtained for a reversible denaturation of the enzyme at pH 12.0 and 2.0. The K-m for N-alpha-benzoyl-L-arginine ethyl ester at pH 8.0 in 20 mM CaCl2-0.1 M KCl-10 mM Tris-HCl buffer at 30 degrees was found to be 7.7 plus or minus 1.9 times 10-6 M and the esterase activity was observed to be dependent on an ionizing group with pK-a equals 5.85. In 2H2O this pKa was increased to 6.35 and the rate of hydrolysis dicreased threefold. The rate of hydrolysis was independent of pH between 8 and 10. The inhibition of the enzyme with L-1-chloro-3-tosylamido-4-phenyl-2-butanone was shown to be associated with the alkylation of its single histidine residue. This residue is present in a homologous amino acid sequence as the active-site histidine in trypsin and chymotrypsin. Optical rotatory dispersion and circular dichroism measurements over the pH range 5.3-10.5 indicated no significant conformational change until the pH was increased above 10.1. The observation that, under the conditions tested, acetylation and carbamylation of the NH2-terminal valine were incomplete is consistent with the view that this group is buried as an ion pair and only becomes available for deprotonation and reaction upon denaturation of the enzyme at pH values greater than 10.0.  相似文献   

10.
G R Parr  G G Hammes 《Biochemistry》1976,15(4):857-862
The kinetics of dissociation and reassembly of rabbit skeletal muscle phosphofructokinase has been studied using fluorescence, stopped-flow fluorescence and enzyme activity measurements. The dissociation of the fully active tetramer in 0.8 M guanidine hydrochloride (0.1 M potassium phosphate, pH 8.0) occurs in three kinetic phases as measured by changes in the protein fluorescence emission intensity: dissociation of tetramer to dimer with a relaxation time of a few milliseconds; dissociation of dimer to monomer with a relaxation time of a few seconds; and a conformational change of the monomer with a relaxation time of a few minutes. All three phases exhibit first-order kinetics; ATP (0.05 mM) retards the second step but does not influence the rate of the other two processes. The rate of the second process increases with decreasing temperature; this may be due to the involvement of hydrophobic interactions in the stabilization of the dimeric enzyme. A further unfolding of the monomer polypeptide chain occurs at higher guanidine concentrations, and the relaxation time associated with this process was found to be 83 ms in 2.5 M guanidine, 0.1 M potassium phosphate (pH 8.0) at 23 degrees C. The phosphofructokinase monomers were reassembled from 0.8 M guanidine chloride by 1:10 dilution of the guanidine hydrochloride concentration and yielded a protein with 70-94% of the original activity, depending on the protein concentration. The reactivation process follows second-order kinetics; ATP (5 mM) increases the rate of reactivation without altering the reaction order, while fructose 6-phosphate does not influence the rate of reaction. The rate-determining step is probably the association of monomers to form the dimer.  相似文献   

11.
The effects of urea and several methylamine solutes on the catalytic stability and aggregation properties of rabbit muscle phosphofructokinase were assessed at physiologically realistic concentrations of the solutes under several pH and temperature regimes. The loss of catalytic activity observed under conditions of pH-induced cold lability was significantly reduced in the presence of trimethylamine-N-oxide, N-trimethylglycine and N-methylglycine (order of decreasing effectiveness). The concentration-dependent methylamine stabilization of the enzyme, seen with as little as 50 mM trimethylamine-N-oxide, was accompanied by increased aggregation of the enzyme to molecular weights greater than the tetramer (polytetramer) as solute concentration was raised to 400 mM. At pH 6.5-6.7 and 25 degrees C, concentrations of urea greater than 25 mM promoted a time-dependent inactivation of the enzyme which was enhanced at lower temperatures. The urea sensitivity of the enzyme exhibited with 0.8 M urea for 1 h at pH 8.0 did not result in measurable inactivation. The fluorescence emission wavelength maximum of the enzyme was shifted to longer wavelengths and the fluorescence intensity was increased as pH was lowered to 7.0, suggesting the occurrence of a protein conformation change as specific amino acid residues of the tetramer became protonated. Measurements of enzyme light scattering indicated that perturbation by urea was correlated with tetramer dissociation, which was irreversible by dialysis at 25 degrees C. The urea and methylamine influences on phosphofructokinase activity and structure were not counteracting. The synergistic interactions among pH, temperature, and solutes observed with phosphofructokinase are compared to effects on other associating-dissociating protein systems in order to evaluate possible mechanisms of action of these low molecular weight solutes.  相似文献   

12.
The effect of temperature on the activity and structural stability of an acid phosphatase (EC 3.1.3.2.) purified from castor bean (Ricinus communis L.) seeds have been examined. The enzyme showed high activity at 45 degrees C using p-nitrophenylphosphate (p-NPP) as substrate. The activation energy for the catalyzed reaction was 55.2 kJ mol(-1) and the enzyme maintained 50% of its activity even after 30 min at 55 degrees C. Thermal inactivation studies showed an influence of pH in the loss of enzymatic activity at 60 degrees C. A noticeable protective effect from thermal inactivation was observed when the enzyme was preincubated, at 60 degrees C, with the reaction products inorganic phosphate-P (10 mM) and p-nitrophenol-p-NP(10 mM). Denaturation studies showed a relatively high transition temperature (Tm) value of 75 degrees C and an influence of the combination of Pi (10 mM) and p-NP (10 mM) was observed on the conformational behaviour of the macromolecule.  相似文献   

13.
Protease secreted into the culture medium by alkalophilic Thermoactinomyces sp. HS682 was purified to an electrophoretically homogeneous state through only two chromatographies using Butyl-Toyopearl 650M and SP-Toyopearl 650S columns. The purified enzyme has an apparent relative molecular mass of 25,000 according to gel filtration on a Sephadex G-75 column and SDS-PAGE and an isoelectric point above 11.0. Its proteolytic activity was inhibited by active-site inhibitors of serine protease, DFP and PMSF, and metal ions, Cu2+ and Hg2+. The enzyme was stable toward some detergents, sodium perborate, sodium triphosphate, sodium-n-dodecylbenzenesulfonate, and sodium dodecyl sulfate, at a concentration of 0.1% and pH 11.5 and 37 degrees C for 60 min. The optimum pH was pH 11.5-13.0 at 37 degrees C and the optimum temperature was 70 degrees C at pH 11.5. Calcium divalent cation raised the pH and heat stabilities of the enzyme. In the presence of 5 mM CaCl2, it showed maximum proteolytic activity at 80 degrees C and stability from pH 4-12.5 at 60 degrees C and below 75 degrees C at pH 11.5. The stabilization by Ca2+ was observed in secondary conformation deduced from the circular dichroic spectrum of the enzyme. The protease hydrolyzed the ester bond of benzoyl leucine ester well. The amino acid terminal sequence of the enzyme showed high homology with those of microbial serine protease, although alanine of the NH2-terminal amino acid was deleted.  相似文献   

14.
An alkaline protease was isolated from culture filtrate of B. subtilis NCIM 2713 by ammonium sulphate precipitation and was purified by gel filtration. With casein as a substrate, the proteolytic activity of the purified protease was found to be optimal at pH 8.0 and temperature 70 degrees C. The purified protease had molecular weight 20 kDa, Isoelectric point 5.2 and km 2.5 mg ml(-1). The enzyme was stable over the pH range 6.5-9.0 at 37 degrees C for 3 hr. During chromatographic separation this protease was found to be susceptible to autolytic degradation in the absence of Ca2+. Ca2+ was not only required for the enzyme activity but also for the stability of the enzyme above 50 degrees C. About 62% activity was retained after 60 min at pH 8.0 and 55 degrees C. DFP and PMSF completely inhibited the activity of this enzyme, while in the presence of EDTA only 33% activity remained. However, it was not affected either by sulfhydryl reagent, or by divalent metal cations, except SDS and Hg2+. The results indicated that this is a serine protease.  相似文献   

15.
The effect of ions on the thermostability and unfolding of Na,K-ATPase from shark salt gland was studied and compared with that of Na,K-ATPase from pig kidney by using differential scanning calorimetry (DSC) and activity assays. In 1 mM histidine at pH 7, the shark enzyme inactivates rapidly at 20 degrees C, as does the kidney enzyme at 42 degrees C (but not at 20 degrees C). Increasing ionic strength by addition of 20 mM histidine, or of 1 mM NaCl or KCl, protects both enzymes against this rapid inactivation. As detected by DSC, the shark enzyme undergoes thermal unfolding at lower temperature (Tm approximately 45 degrees C) than does the kidney enzyme (Tm approximately 55 degrees C). Both calorimetric endotherms indicate multi-step unfolding, probably associated with different cooperative domains. Whereas the overall heat of unfolding is similar for the kidney enzyme in either 1 mM or 20 mM histidine, components with high mid-point temperatures are lost from the unfolding transition of the shark enzyme in 1 mM histidine, relative to that in 20 mM histidine. This is attributed to partial unfolding of the enzyme due to a high hydrostatic pressure during centrifugation of DSC samples at low ionic strength, which correlates with inactivation measurements. Addition of 10 mM NaCl to shark enzyme in 1 mM histidine protects against inactivation during centrifugation of the DSC sample, but incubation for 1 h at 20 degrees C prior to addition of NaCl results in loss of components with lower mid-point temperatures within the unfolding transition. Cations at millimolar concentration therefore afford at least two distinct modes of stabilization, likely affecting separate cooperative domains. The different thermal stabilities and denaturation temperatures of the two Na,K-ATPases correlate with the respective physiological temperatures, and may be attributed to the different lipid environments.  相似文献   

16.
AIMS: Isolation and screening of extreme halophilic archaeon producing extracellular haloalkaliphilic protease and optimization of culture conditions for its maximum production. METHODS AND RESULTS: Halogeometricum sp. TSS101 was isolated from salt samples and screened for the secretion of protease on gelatin and casein plates containing 20% NaCl. The archaeon was grown aerobically in a 250 ml flask containing 50 ml of (w/v) NaCl 20%; MgCl(2) 1%; KCl 0.5%; trisodium citrate 0.3%; and peptone 1%; pH 7.2 at 40 degrees C on rotary shaker. The production of enzyme was investigated at various pH, temperatures, NaCl concentrations, metal ions and different carbon and nitrogen sources. The partially purified protease had activity in a broad pH range (7.0-10.0) with optimum activity at pH 10.0 and a temperature (60 degrees C). The enzyme was thermostable and retained 70% initial activity at 80 degrees C. Maximum protease production occurred at 40 degrees C in a medium containing 20% NaCl (w/v) and 1% skim milk powder after 84 h in shaking culture. Enzyme secretion was observed at a broad pH range of 7.0-10.0. Addition of CaCl(2) (200 mmol) to the culture medium enhanced the production of protease. Protein rich flours proved to be cheap and good alternative source for enzyme production. Different osmolytes were tested for the growth and production of haloalkaliphilc protease and found that betaine and glycerol enhanced growth without secretion of the protease. Immobilization studies showed that whole cells immobilized in 2% alginate beads were stable up to 10 batches and able to secrete the protease, which attained maximum production within 60 h under shaking conditions. CONCLUSIONS: Halogeometricum sp. TSS101 secreted an extracellular haloalkaliphilic and thermostable protease. The optimum conditions required for maximum production are 20% NaCl, 1% skim milk powder and temperature at 40 degrees C. Addition of CaCl(2) (200 mmol) enhanced the enzyme production. Immobilization of whole cells in absence of NaCl proved to be useful for continuous production of haloalkaliphilic protease. SIGNIFICANCE AND IMPACT OF THE STudy: The low cost protein rich flours were used as an alternative carbon and nitrogen sources for enzyme production. Immobilization of halophilic cells in alginate beads can be used in continuous production of halophilic enzyme. The halophilic and thermostable protease from Halogeometricum sp. TSS101 is good source for industrial applications and can be a suitable source for preparation of fish sauce.  相似文献   

17.
A new extracellular protease having a prospective application in the food industry was isolated from Bacillus sUbtilis NCIM 2711 by (NH4)2SO4 precipitation from the cell broth. It was purified using DEAE-Cellulose and CM-Sephadex C-50 ion-exchange chromatography. With casein as a substrate, the proteolytic activity of the purified protease was found to be optimal at pH 7.0 and temperature 55 degrees C with Km 1.06 mg/ml. The enzyme was stable over a pH range 6.5-8.0 at 30 degrees C for 1 hr in presence of CaCl2 x 2H2O. At 55 degrees C, the enzyme retained 60% activity up to 15 min in presence of CaCl2 x 2H2O. EDTA and o-phenanthroline (OP) completely inhibited the enzyme activity while DFP, PMSF and iodoacetamide were ineffective. The enzyme was completely inhibited by Hg2+ and partially by Cd2+, Cu2+, Ni2+, Pb2+ and Fe2+. The OP inhibited enzyme could be reactivated by Zn2+ and Co2+ up to 75% and 69% respectively. It is a neutral metalloprotease showing a single band of 43 kDa on SDS-PAGE.  相似文献   

18.
D E Hill  G G Hammes 《Biochemistry》1975,14(2):203-213
Equilibrium binding studies of the interaction of rabbit muscle phosphofructokinase with fructose 6-phosphate and fructose 1,6-bisphosphate have been carried out at 5 degrees in the presence of 1-10 mM potassium phosphate (pH 7.0 and 8.0), 5 mM citrate (pH 7.0), or 0.22 mm adenylyl imidodiphosphate (pH 7.0 and 8.0). The binding isotherms for both fructose 6-phosphate and fructose 1,6-bisphosphate exhibit negative cooperativity at pH 7.0 and 8.0 in the presence of 1-10 mM potassium phosphate at protein concentrations where the enzyme exists as a mixture of dimers and tetramers (pH 7.0) or as tetramers (pH 8.0) and at pH 7.0 in the presence of 5 mM citrate where the enzyme exists primarily as dimers. The enzyme binds 1 mol of either fructose phosphate/mol of enzyme monomer (molecular weight 80,000). When enzyme aggregation states smaller than the tetramer are present, the saturation of the enzyme with either ligand is paralleled by polymerization of the enzyme to tetramer, by an increase in enzymatic activity and by a quenching of the protein fluorescence. At protein concentrations where aggregates higher than the tetramer predominate, the fructose 1,6-bisphosphate binding isotherms are hyperbolic. These results can be quantitatively analyzed in terms of a model in which the dimer is associated with extreme negative cooperativity in binding the ligands, the tetramer is associated with less negative cooperativity, and aggregates larger than the tetramer are associated with little or no cooperativity in the binding process. Phosphate is a competitive inhibitor of the fructose phosphate sites at both pH 7.0 and 8.0, while citrate inhibits binding in a complex, noncompetitive manner. In the presence of the ATP analog adenylyl imidodiphosphate, the enzyme-fructose 6-phosphate binding isotherm is sigmoidal at pH 7.0, but hyperbolic at pH 8.0. The characteristic sigmoidal initial velocity-fructose 6-phosphate isotherms for phosphofructokinase at pH 7.0, therefore, are due to an heterotropic interaction between ATP and fructose 6-phosphate binding sites which alters the homotropic interactions between fructose 6-phosphate binding sites. Thus the homotropic interactions between fructose 6-phosphate binding sites can give rise to positive, negative, or no cooperativity depending upon the pH, the aggregation state of the protein, and the metabolic effectors present. The available data suggest the regulation of phosphofructokinase involves a complex interplay between protein polymerization and homotropic and heterotropic interactions between ligand binding sites.  相似文献   

19.
A tissue carboxypeptidase-A-like enzyme was purified to apparent homogeneity from terminally differentiated epidermal cells of 2-day-old rats by potato inhibitor affinity chromatography followed by FPLC Mono Q column chromatography. The enzyme has an Mr of 35,000 as determined by SDS-polyacrylamide gel electrophoresis and HPLC gel filtration. It has a pH optimum of 8.5 for hydrolysis of benzyloxycarbonyl-Phe-Leu (Km = 0.22 mM, kcat = 57.9 s-1). The enzyme does not hydrolyze substrates with Arg, Lys and Pro at the C-terminal and Pro at the penultimate position. Angiotensin I was effectively hydrolyzed (Km = 0.06 mM, kcat = 6.48 s-1) and produced both des-Leu10-angiotensin I and angiotensin II. The enzyme activity, relatively stable at 4 degrees C and pH 8.0-10.5, was inactivated at pH values higher than 12.0 and lower than 5.0 or at 65 degrees C for 10 min. Inhibitor profiles of the epidermal enzyme also differed slightly from those of tissue carboxypeptidase A of pancreatic or mast cell origin.  相似文献   

20.
Differential scanning calorimetry of Cd(II) alkaline phosphatases   总被引:1,自引:0,他引:1  
Differential scanning calorimetry has been employed to monitor structural alterations induced in the dimeric enzyme alkaline phosphatase on binding of Cd(II) (to the metal-free apoenzyme) and phosphate (Pi) (to the Cd(II) enzyme). Cd(II) addition to the apoenzyme at pH 6.5 results in an increased transition temperature, suggesting a stabilizing effect of the bound metal ion. Two distinct structural forms of the protein are detected as discrete calorimetric transitions (Tm = 69-84 degrees C; 87-94 degrees C, respectively). Distribution of the enzyme between these forms is found to depend on the exogenous Cd(II) concentration and the protocol of Cd(II) addition. These results indicate that conversion between the conformational forms is a slow process which appears to require specific levels of metal ion site occupancy. These studies, in which the exogenous Cd(II) concentration was varied from 10(-5) M to 10(-3) M suggest a structural basis for previously observed hysteretic phenomena observed on Cd(II) binding to the enzyme. Even at a minimum stoichiometry of Cd(II) (2 eq/mol of dimer) a single equivalent of Pi is sufficient to accelerate assumption of a stabilized form of the protein (Tm = 90 degrees C). This is followed by a slow structural change paralleling the time course of formation of the functional 2 Cd(II) phosphoryl enzyme which displays two calorimetric transitions (Tm = 65 degrees C, 88 degrees C). The low temperature transition does not appear if Pi is initially present at millimolar concentrations and is abolished on addition of Pi at concentrations in excess of 0.1 mM. These observations suggest the presence of a second, distinct Pi binding site on the 2 Cd(II) phosphoryl enzyme. This is supported by the changes observed in the 31P NMR chemical shift of Pi added to comparable enzyme samples. These data, including assessment of the effect of the presence of Mg(II), are discussed in terms of the mechanism of metal ion association to the enzyme and rearrangement of bound metal ions induced by Pi binding.  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