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.
Ribonucleoside-diphosphate reductase (EC 1.17.4.1) from Escherichia coli consists of two nonidentical subunits, proteins R1 and R2. The binding of the product dCDP to protein R1 and to the holoenzyme R1R2 has been studied by means of 1H-NMR spectroscopy. In presence of the effector dTTP at 25 degrees C, dCDP was found to be in rapid exchange between the binding sites and the solvent which results in a broadening of the dCDP resonances. When both proteins R1 and R2 are present, so that the complex R1R2 is formed, a smaller broadening is observed than with protein R1 alone. No further linewidth decrease was observed when the [R2]/[R1] ratio exceeded 1. The binding constant of dCDP to R1 or R1R2 is the same, Kd = 0.9 mM. The smaller broadening of the dCDP resonances observed with the complex R1R2 as compared with R1 may be explained by the combination of two effects: (a) the overall tumbling time of the protein will increase when going from R1 to R1R2, which will cause the broadening to increase correspondingly, and (b) a twofold decrease of the number of binding sites in rapid exchange, which will decrease the broadening by a factor of 0.5. The effect of R2 without iron (apoR2) is reduced compared with native R2, probably because of some denatured proteins, while a C-terminal peptide from R2 did not cause any narrowing at all.  相似文献   

2.
Nuclear magnetic relaxation methods were used to investigate the interaction of the inhibitor succinate with aspartate transcarbamylase from Escherichia coli. Over the pH range 7 to 9, the dissociation constant for succinate remains less than the inhibitor concentration used for most of this work (0.05 M). As a result, the enzyme predominantly exists in a single "gross" conformational state. Succinate binding to this enzyme state (generally known as the R form) parallels the behavior seen previously with the isolated catalytic subunit (Beard, C. B., and Schmidt, P.G. (1973) Biochemistry 12, 2255-2264). The pH and temperature dependence of succinate proton relaxation rates, 1/T2 - 1/T1, in the presence of carbamyl phosphate, is interpreted in terms of a binding mechanism involving three forms of the enzyme, differing by their states of protonation. The least protonated form of the enzyme does not interact with succinate, the singly protonated species binds succinate to form a rapidly dissociating complex, and the doubly protonated species undergoes a conformational isomerization upon succinate binding, yielding a slow exchange complex. Relaxation data provide sufficient information to determine pKa values of 7.2 and 8.9 for two ionizing groups, as well as the dissociation constant for succinate in the fast exchange complex, Kd =1.6 X 10(-2) M. Rate constants for the forward and reverse steps of the isomerization, 1.3 X 10(3) s-1 and 33 s-1, respectively, indicate a significantly slower reverse rate from that obtained in the earlier NMR study of the isolated catalytic subunit. In experiments where the succinate concentration was varied, the relaxation rates showed sigmoidal binding of that ligand to the fast exchange complex above pH 9.1, (a) indicating cooperative binding of succinate, and (b) suggesting that above pH 9.1, the system cannot be characterized by a single dissociation constant, ionization constant, or relaxation effect. CTP and ATP were tested for their ability to affect succinate binding to the fast exchange complex. Heterotropic interactions were observed for CTP but not for ATP. Addition of low concentrations of the transition state analog N-(phosphonacetyl)-L-aspartate to the enzyme-carbamyl phosphate-succinate complex sharply decreased the relaxation rate, indicating that the measurements are sensitive only to succinate bound specifically to the active site.  相似文献   

3.
The reaction of choline acetyltransferase with methoxycarbonyl alkyl disulfides leads to a progressive loss in enzyme activity as the size of the alkyl group increases from methyl to n-butyl. Reaction with 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) or methoxycarbonyl coenzyme A (CoA) disulfide, leads to a total loss of enzyme activity. DTNB inactivation is biphasic (k1 = approximately 9 x 10(2) M-1 s-1, k2 = approximately 6 x 10(1) M-1 s-1) with the slow phase being diminished by acetyl-CoA. Methoxycarbonyl-CoA disulfide inactivation is also biphasic (k1 = approximately 2.1 x 10(3) M-1 s-1, k2 = approximately 6 x 10(1) M-1 s-1), with the rapid phase being diminished in the presence of acetyl-CoA. Inactivation by methoxycarbonyl methyl disulfide, ethyl disulfide, or hydroxyethyl disulfide, or by methyl methanethiosulfonate is not biphasic. Pretreatment of the enzyme with methyl methanethiosulfonate, which leads to a 25% loss in enzyme activity, abolishes the fast phase of DTNB inactivation, the slow phase of methoxycarbonyl-CoA disulfide inactivation, and any further inactivation by methoxycarbonyl ethyl disulfide. These results are interpreted to suggest that choline acetyltransferase contains two classes of reactive sulfhydryl groups, neither of which are required for enzyme activity.  相似文献   

4.
Steady-state electron flow through and electron delivery into isolated dimeric bc1 complex (ubiquinol--cytochrome c oxidoreductase) from Neurospora crassa and beef heart mitochondria were studied in the presence of increasing concentrations of antimycin A, funiculosin and/or myxothiazol. Parabolic or linear inhibition curves were obtained, depending upon the different quinols and inhibitors that were used. Linear curves occur when the inhibitor directly affects the rate-determining step. The most reasonable explanation for the parabolic curves is given by a fast intradimeric exchange of the hydrophobic inhibitors antimycin A, funiculosin (rate less than 500 s-1) and of myxothiazol (rate greater than 1 s-1). Using mitochondria from beef heart, the shape of the inhibition curve with antimycin A is parabolic if the quinol--O2 oxidoreductase turns over at about 300 s-1, but hyperbolic if the rate is 5 times less. The hyperbolic titration curve may be the result of both intradimeric and an additional interdimeric redistribution (rate approximately 100 s-1) of inhibitors between enzymes incorporated in a continuous phospholipid membrane. This explanation is supported by experiments with chromatophores obtained from Rhodobacter capsulatus. As recently described [Fernandez-Velasco, J. & Crofts, A. R. (1992) Biophys. J. 2, A153], cytochrome b becomes fully reoxidized within 1 s after a flash at substoichiometric concentrations of antimycin A. This kinetic of the slow reoxidation can be expressed in terms of the intradimeric and interdimeric redistribution with rate constants of about 10 s-1 and 2 x 10(6) M-1 s-1, respectively. It seems that rapid inhibitor redistribution may be a widespread phenomenon for hydrophobic inhibitors of enzymes incorporated in lipid membranes.  相似文献   

5.
Acryloyl-CoA reductase from Clostridium propionicum catalyses the irreversible NADH-dependent formation of propionyl-CoA from acryloyl-CoA. Purification yielded a heterohexadecameric yellow-greenish enzyme complex [(alpha2betagamma)4; molecular mass 600 +/- 50 kDa] composed of a propionyl-CoA dehydrogenase (alpha2, 2 x 40 kDa) and an electron-transferring flavoprotein (ETF; beta, 38 kDa; gamma, 29 kDa). A flavin content (90% FAD and 10% FMN) of 2.4 mol per alpha2betagamma subcomplex (149 kDa) was determined. A substrate alternative to acryloyl-CoA (Km = 2 +/- 1 microm; kcat = 4.5 s-1 at 100 microm NADH) is 3-buten-2-one (methyl vinyl ketone; Km = 1800 microm; kcat = 29 s-1 at 300 microm NADH). The enzyme complex exhibits acyl-CoA dehydrogenase activity with propionyl-CoA (Km = 50 microm; kcat = 2.0 s-1) or butyryl-CoA (Km = 100 microm; kcat = 3.5 s-1) as electron donor and 200 microm ferricenium hexafluorophosphate as acceptor. The enzyme also catalysed the oxidation of NADH by iodonitrosotetrazolium chloride (diaphorase activity) or by air, which led to the formation of H2O2 (NADH oxidase activity). The N-terminus of the dimeric propionyl-CoA dehydrogenase subunit is similar to those of butyryl-CoA dehydrogenases from several clostridia and related anaerobes (up to 55% sequence identity). The N-termini of the beta and gamma subunits share 40% and 35% sequence identities with those of the A and B subunits of the ETF from Megasphaera elsdenii, respectively, and up to 60% with those of putative ETFs from other anaerobes. Acryloyl-CoA reductase from C. propionicum has been characterized as a soluble enzyme, with kinetic properties perfectly adapted to the requirements of the organism. The enzyme appears not to be involved in anaerobic respiration with NADH or reduced ferredoxin as electron donors. There is no relationship to the trans-2-enoyl-CoA reductases from various organisms or the recently described acryloyl-CoA reductase activity of propionyl-CoA synthase from Chloroflexus aurantiacus.  相似文献   

6.
Dynamic NMR methods, such as differential line broadening and transferred NOE spectroscopy, are normally reserved for the study of small molecule ligand interactions with large protein receptors. Using a combination of isotope labeling and isotope edited NMR, we have extended these techniques to characterize interactions of a much larger protein/drug complex, FKBP-12/ FK506 with its receptor protein, calcineurin. In order to examine this multicomponent system by dynamic NMR methods, the 93 kDa, tightly bound FKBP-12/FK506/Cn complex was replaced with a lower affinity, rapidly exchanging system consisting of FKBP-12/FK506 (13 kDa), recombinant calcineurin subunit B (CnB) (20 kDa), and a synthetic peptide (4 kDa) corresponding to the B binding domain (BBD) of calcineurin catalytic subunit A (CnA). Analysis of 1H-13C HSQC data acquired for the FKBP-12/ 13C-FK506 and FKBP-12/13C-FK506/CnB/BBD complexes indicates that FKBP-12/FK506 and CnB/BBD are in fast exchange in the quaternary complex. Comparison of proton line widths shows significant broadening of resonances along the macrocycle backbone at 13-CH, 13-OMe, 15-OMe, 18-CH2, 20-CH, 21-CH, and 25-Me, as well as moderate broadening on the macrocycle backbone at 17-Me, 24-CH, and the pyranose 12-CH2 protons. The tri-substituted olefin and cyclohexyl groups also show moderate broadening at the 27-Me, 28-CH, and 30-CH2 positions, respectively. Unexpectedly, little line broadening was observed for the allyl resonances of FK506 in the quaternary complex, although 13C longitudinal relaxation measurements suggest this group also makes contacts with calcineurin. In addition, intermolecular transfer NOE peaks were observed for the allyl 37-CH2, 21-CH, 30-CH2, 13-OMe, 15-OMe, 17-Me, 25-Me, and 27-Me groups, indicating that these are potential sites on the FK506 molecule that interact with calcineurin.  相似文献   

7.
A pyrimidine-guanine sequence-specific ribonuclease (RC-RNase) was purified from Rana catesbeiana (bullfrog) oocytes by sequential phosphocellulose, Sephadex G75, heparin Sepharose CL 6B and CM-Sepharose CL 6B column chromatography. The purified enzyme with molecular weight of 13,000 daltons gave a single band on SDS-polyacrylamide gel. One CNBr-cleaved fragment has a sequence of NVLSTTRFQLNT/TRTSITPR, which is identical to residues 59-79 of a sialic acid binding lectin from R. catesbeiana eggs, and is 71% homologous to residues 60-80 of an RNase from R. catesbeaina liver. The RC-RNase preferentially cleaved RNA at pyrimidine residues with a 3' flanking guanine under various conditions. The sequence specificity of RC-RNase was further confirmed with dinucleotide as substrates, which were analyzed by thin layer chromatography after enzyme digestion. The values of kcat/km for pCpG, pUpG and pUpU were 2.66 x 10(7) M-1s-1, 2.50 x 10(7) M-1s-1 and 2.44 x 10(6) M-1s-1 respectively, however, those for other phosphorylated dinucleotides were less than 2% of pCpG and pUpG. As compared to single strand RNA, double strand RNA was relatively resistant to RC-RNase. Besides poly (A) and poly (G), most of synthetic homo- and heteropolynucleotides were also susceptible to RC-RNase. The RC-RNase was stable in the acidic (pH 2) and alkaline (pH 12) condition, but could be inactivated by heating to 80 degrees C for 15 min. No divalent cation was required for its activity. Furthermore, the enzyme activity could be enhanced by 2 M urea, and inhibited to 50% by 0.12 M NaCl or 0.02% SDS.  相似文献   

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

9.
1. The kinetics of the interaction of cytochrome c2 and photosynthetic reaction centers purified from Rhodobacter capsulatus were studied in proteoliposomes reconstituted with a mixture of phospholipids simulating the native membrane (i.e. containing 25% L-alpha-phosphatidylglycerol). 2. At low ionic strength, the kinetics of cytochrome-c2 oxidation induced by a single turnover flash was very different, depending on the concentration of cytochrome c2: at concentrations lower than 1 microM, the process was strictly bimolecular (second-order rate constant, k = 1.7 x 10(9) M-1 s-1), while at higher concentrations a fast oxidation process (half-time lower than 20 microseconds) became increasingly dominant and encompassed the total process at a cytochrome c2 concentration around 10 microM. From the concentration dependence of the amplitude of this fast phase an association constant for a reaction-center--cytochrome-c2 complex of about 10(5) M-1 was evaluated. From the fraction of photo-oxidized reaction centers promptly re-reduced in the presence of saturating concentrations of externally added cytochrome c2, it was found that in approximately 60% of the centers the cytochrome-c2 site was exposed to the external compartment. 3. Both the second-order oxidation reaction and the formation of the reaction-center--cytochrome-c2 complex were very sensitive to ionic strength. In the presence of 180 mM KCl, the value of the second-order rate constant was decreased to 7.0 x 10(7) M-1 s-1 and no fast oxidation of cytochrome c2 could be observed at 10 microM cytochrome c2. 4. The kinetics of exchange of oxidized cytochrome c2 bound to the reaction center with the reduced form of the same carrier, following a single turnover flash, was studied in double-flash experiments, varying the dark time between photoactivations over the range 30 microseconds to 5ms. The experimental results were analyzed according to aminimal kinetic model relating the amounts of oxidized cytochrome c2 and reaction centers observable after the second flash to the dark time between flashes. This model included the rate constants for the electron transfer between the primary and secondary ubiquinone acceptors of the complex (k1) and for the exchange of cytochrome c2 (k2). Fitting to the experimental results indicated a value of k1 equal to 2.4 x 10(3) s-1 and a lower limit for k2 of approximately 2 x 10(4) s-1 (corresponding to a second-order rate constant of approximately 3 x 10(9) M-1 s-1).  相似文献   

10.
The molecular mass of destabilase isolated from the medicinae leech Hirudo medicinalis was found to be equal to 12.3 kDa. A kinetic analysis of the sole presently known synthetic substrate, L-gamma-Glu-pNA, showed that the enzyme is relatively stable to heating (5 min, 70 degrees C); the pH optimum lies at 7.0-8.5. The enzyme has a specific activity of 0.15 x 10(-9) mol.s-1.mg-1; Km = 2.2 x 10(-4) M, kcat is 3.53 x 10(-3) s-1 (pH 8.0, 37 degrees C).  相似文献   

11.
The time course of the interaction between trypsin and a synthetic peptide corresponding to a segment (residues 676-703) of the bait region (residues 666-706) of human alpha 2-macroglobulin (alpha 2M) was studied by measuring the generation of cleavage products as a function of time by HPLC. Three primary cleavage sites for trypsin were present in the synthetic peptide. The fastest cleavage occurred at the bond corresponding to Arg696-Leu in alpha 2M with an estimated kcat/Km = 1-2 x 10(6) M-1.s-1. This value is of the same magnitude as that characterizing the interaction of alpha 2M and trypsin when taking into account the fact that alpha 2M is a tetramer, kcat/Km = 5 x 10(6) M-1.s-1 [Christensen, U. & Sottrup-Jensen, L. (1984) Biochemistry 23, 6619-6626]. The values of kcat/Km for cleavage at bonds corresponding to Arg681-Val and Arg692-Gly in alpha 2M were 1.5 x 10(5) M-1.s-1 and 1.3 x 10(5) M-1.s-1, respectively. Cleavage of intermediate product peptides was slower, with kcat/Km in the range 13-1.3 x 10(6) M-1.s-1. The value of Km determined for fast cleavage in the synthetic peptide was 8-10 microM. 1H-NMR spectroscopy indicated no ordered structure of the peptide. Hence, the very fast cleavage of the peptide is compatible with a loose structure that readily adopts a conformation favorable for recognition and cleavage by trypsin.  相似文献   

12.
Glycolate oxidase from spinach has been expressed in Saccharomyces cerevisiae. The active enzyme was purified to near-homogeneity (purification factor approximately 1400-fold) by means of hydroxyapatite and anion-exchange chromatography. The purified glycolate oxidase is nonfluorescent and has absorbance peaks at 448 (epsilon = 9200 M-1 cm-1) and 346 nm in 0.1 M phosphate buffer, pH 8.3. The large bathochromic shift of the near-UV band indicates that the N(3) position is deprotonated at pH 8.3. A pH titration revealed that the pK of the N(3) is shifted from 10.3 in free flavin to 6.4 in glycolate oxidase. Glycolate oxidase is competitively inhibited by oxalate with a Kd of 0.24 mM at 4 degrees C in 0.1 M phosphate buffer, pH 8.3. Three pieces of evidence demonstrate that glycolate oxidase stabilizes a negative charge at the N(1)-C(2 = O) locus: the enzyme forms a tight sulfite complex with a Kd of 2.7 x 10(-7) M and stabilizes the anionic flavosemiquinone and the benzoquinoid form of 8-mercapto-FMN. Steady-state analysis at pH 8.3, 4 degrees C, yielded a Km = 1 x 10(-3) M for glycolate and Km = 2.1 x 10(-4) M for oxygen. The turnover number has been determined to be 20 s-1. Stopped-flow studies of the reductive (k = 25 s-1) and oxidative (k = 8.5 x 10(4) M-1 s-1) half-reactions have identified the reduction of glycolate oxidase to be the rate-limiting step.  相似文献   

13.
Carbon monoxide chlorocruorin from Eudistylia vancouverii shows three distinct first-order relaxations with rates of 2.9 x 10(9) s-1, 6.5 x 10(7) s-1, and 3.2 x 10(6) s-1 (geminate reactions) and three second-order relaxations with rates of 4.7 x 10(6) M-1 s-1, 7 x 10(5) M-1 s-1, and 7 x 10(4) M-1 s-1, when studied by flash photolysis. The amplitudes of the second-order reactions depend on the extent of photolysis. This may be due to relaxation from the liganded (R) to the unliganded (T) conformation following photolysis and suggests that the combination rates contribute to cooperativity. In a stopped-flow experiment only the slowest phase with a rate of 7 x 10(4) M-1 s-1 is observed. It is assigned to binding to the T-state protein. Fragments of the native protein containing 12 and 4 hemes react like the holoprotein suggesting that the tetramer is a major cooperative unit. Oxygen binding shows three geminate relaxations with rates of 2.5 x 10(10) s-1, 3.5 x 10(7) s-1, and 4.5 x 10(6) s-1, and two second-order rates of 1.5 x 10(7) M-1 s-1 and 1 x 10(6) M-1 s-1. The amplitudes of the second-order phases do not correlate with the extent of photolysis. The results with the two ligands are consistent with an allosteric transition fast enough to compete with a rebinding rate of 500 s-1 in the R to T direction (CO rebinding) but not fast enough to compete with oxygen rebinding. There is significant heterogeneity in the R-state kinetics, but the T-state reaction is homogeneous.  相似文献   

14.
The electron transfer reaction between ferrocyanide ion and the blue copper protein, stellacyanin, has been investigated by means of 13C NMR line broadening of the inorganic oxidant. The temperature dependence of the ferrocyanide line broadening gives an activation energy for the electron transfer reaction of 17 +/- 3 kJ. The apparent rate constant decreases with increasing concentration of K4Fe(CN)6, a result which can be explained either by formation of a strong precursor ferrocyanide--stellacyanin [Cu(II)] complex or by increased formation of KFe(CN)3-6 ion pairs. The direct electron transfer between ferrocyanide and ferricyanide has also been studied by 13C NMR line broadening of the former species. The ferricyanide concentration dependence of the exchange line broadening yields a value for the apparent second-order rate constant at 25 degrees C of k = 1.65 . 10(3) M-1 . s-1, in agreement with previously reported values derived from 14N NMR and isotope exchange studies. This rate constant shows a linear dependence on the K+ concentration, independent of ionic strength, a result which confirms the importance of ion pair species such as KFe(CN)3-6 and KFe(CN)2-6 in the direct electron transfer mechanism. The general applications of the method are discussed, including the considerations which suggest that a wide range of electron transfer rates, from about 1 s-1 to 4 . 10(3) s-1, are, in principle, accessible to this technique. The potential utility of ferrocyanide 13C spin--lattice relaxation time measurements is decreasing the lower limit of this range is also discussed.  相似文献   

15.
K C Cho  J J Hopfield 《Biochemistry》1979,18(26):5826-5833
The molecular mechanism of hemoglobin cooperativity was studied kinetically by flash photolysis on mixed-state hemoglobins which consist of three ferrous carboxy subunits and one hybrid ferric subunit including fluoromet, azidomet, cyanatomet, and thiocyanatomet. The effects of conformational transitions on the hybrid subunit were detected by kinetic absorption spectroscopy after the CO was fully photodissociated from the binding sites by a large pulse of light from a tunable dye laser. The hemoglobin conformational transition rate was observed to depend on its state of ligation. At 22 degrees C, pH 7, and 0.1 M phosphate, the deoxy R yields T conformational change rate is 4 x 10(4)s-1. The rate decreases to 1.4 x 10(4)s-1 for singly ligated hemoglobin. The R yields T conformation change alters the energy separation between high- and low-spin states for azidomet, cyanatomet, and thiocyanatomet subunits by about 700, 300, and 300 cal/mol, respectively. There are two possible implications of this result: (1) the iron atom spin state is not the only major factor in the determination of its position with respect to the heme plane or (2) the change with conformation of the protein force exerted by the proximal histidine on the iron atom (for an iron to heme-plane displacement of less than 0.3 A) is less than 50% of that expected from simple models in which this motion is responsible for cooperativity.  相似文献   

16.
A kinetic analysis of MspI DNA methyltransferase (M.MspI) is presented. The enzyme catalyzes methylation of lambda-DNA, a 50-kilobase pair linear molecule with multiple M.MspI-specific sites, with a specificity constant (kcat/KM) of 0.9 x 10(8) M-1 s-1. But the values of the specificity constants for the smaller DNA substrates (121 and 1459 base pairs (bp)) with single methylation target or with multiple targets (sonicated lambda-DNA) were less by an order of magnitude. Product inhibition of the M.MspI-catalyzed methylation reaction by methylated DNA is competitive with respect to DNA and noncompetitive with respect to S-adenosylmethionine (AdoMet). The S-adenosylhomocysteine inhibition of the methylation reaction is competitive with respect to AdoMet and uncompetitive with respect to DNA. The presteady state kinetic analysis showed a burst of product formation when AdoMet was added to the enzyme preincubated with the substrate DNA. The burst is followed by a constant rate of product formation (0.06 mol per mol of enzyme s-1) which is similar to catalytic constants (kcat = approximately 0.056 s-1) measured under steady state conditions. The isotope exchange in chasing the labeled methyltransferase-DNA complex with unlabeled DNA and AdoMet leads to a reduced burst as compared with the one involving chase with labeled DNA and AdoMet. The enzyme is capable of exchanging tritium at C-5 of target cytosine in the substrate DNA in the absence of cofactor AdoMet. The kinetic data are consistent with an ordered Bi Bi mechanism for the M.MspI-catalyzed DNA methylation where DNA binds first.  相似文献   

17.
Urease has been purified from the dehusked seeds of pigeonpea (Cajanus cajan L.) to apparent electrophoretic homogeneity with approximately 200 fold purification, with a specific activity of 6.24 x10(3) U mg(-1) protein. The enzyme was purified by the sequence of steps, namely, first acetone fractionation, acid step, a second acetone fractionation followed by gel filtration and anion-exchange chromatographies. Single band was observed in both native- and SDS-PAGE. The molecular mass estimated for the native enzyme was 540 kDa whereas subunit values of 90 kDa were determined. Hence, urease is a hexamer of identical subunits. Nickel was observed in the purified enzyme from atomic absorption spectroscopy with approximately 2 nickel ions per enzyme subunit. Both jack bean and soybean ureases are serologically related to pigeonpea urease. The amino acid composition of pigeonpea urease shows high acidic amino acid content. The N-terminal sequence of pigeonpea urease, determined up to the 20th residue, was homologous to that of jack bean and soybean seed ureases. The optimum pH was 7.3 in the pH range 5.0-8.5. Pigeonpea urease shows K(m) for urea of 3.0+/-0.2 mM in 0.05 M Tris-acetate buffer, pH 7.3, at 37 degrees C. The turnover number, k(cat), was observed to be 6.2 x 10(4) s(-1) and k(cat)/K(m) was 2.1 x 10(7) M(-1) s(-1). Pigeonpea urease shows high specificity for its primary substrate urea.  相似文献   

18.
We have measured the forward and reverse rates of the allosteric transition of hemoglobin A with three CO molecules bound by using modulated excitation coupled with fluorescence quenching of the DPG analogue, PTS (8-hydroxy-1,3,6 pyrene trisulfonic acid). This dye is observed to bind to the T state with significantly larger affinity than to the R state, and thus provides an unequivocal marker for the molecule's conformational change. The allosteric rates obtained with the fluorescent dye (pH 7.0, bis-Tris buffer) are (3.4 +/- 1.0) x 10(3)s-1 for the R to T transition and (2.1 +/- 0.5) x 10(4)s-1 for the T to R transition. This gives an equilibrium constant L3 of 0.16 +/- 0.06. These results provide good agreement with modulated difference spectra calibrated from model compounds, arguing that there is little if any difference in the kinetics observed by the heme spectra and the kinetics of the full subunit motion. The equilibrium constant between structures (L3) is smaller in the absence of phosphates than observed in phosphate buffer (0.33). However, the rates of the allosteric transition increase in the absence of phosphates as compared with the corresponding rates in phosphate buffer of 1.0 x 10(3)s-1 and 3.0 x 10(3)s-1. The effects of inorganic phosphates on the equilibrium can be separated from the effects on kinetics. We find that phosphates also affect the dynamic behavior of hemoglobin, and the presence of 0.15 M phosphate can be viewed as raising the transition state energy between R and T conformations by approximately 0.5 kcal/mol exclusive of the T state stabilization. Dissociation constants for the dye were measured to be 104 +/- 25 microM for unligated T state and 930 +/- 300 microM for the fully ligated R state. The best fit equilibrium constant (125 +/- 40 microM) for three ligands bound does not differ significantly from that measured without ligands bound. Incidental to the measurement technique is the determination of the rates of binding and release of the dye. The association rate for binding to the T state is large, (at least 4 x 10(9) M-1 s-1) and may be diffusion limited, while the association and dissociation rates for R state binding, while not determined with precision, are clearly much smaller, of the scale of 10(5) M-1 s-1 for association.  相似文献   

19.
The intrinsic reactivity of delta 2- and delta 3-deacetoxy-7-phenylacetamidocephalosporanates, penicillanate, unsubstituted, 2-methyl- and 2-phenyl-penems and other beta-lactam antibiotics has been expressed in terms of the second-order rate constant (M-1.s-1(OH-)) for the hydrolysis of the beta-lactam amide bond by OH- at 37 degrees C. The values thus obtained have been compared with the second-order rate constants (M-1.s-1(enzyme) for the opening of the same beta-lactam amide bond during interaction with the beta-lactamases of Streptomyces albus G and Actinomadura R39 and the D-alanyl-D-alanine-cleaving serine peptidases of Streptomyces R61 and Actinomadura R39. Depending on the cases, the accelerating effect due to enzyme action and expressed by the ratio M-1.s-1(enzyme)/M-1.s-1(OH) varies from less than 2 to more than 1 x 10(6). The primary parameter that governs enzyme action is the goodness of fit of the beta-lactam molecule to the enzyme cavity rather than its intrinsic reactivity. With the D-alanyl-D-alanine-cleaving serine peptidases, the three penems studied form intermediate complexes characterized by very short half lives of 14-100 s, values significantly lower than those exhibited by most beta-lactam compounds.  相似文献   

20.
The "serpin" plasminogen activator inhibitor 1 (PAI-1) is the fast acting inhibitor of plasminogen activators (tissue-type (t-PA) and urokinase type-PA) and is an essential regulatory protein of the fibrinolytic system. Its P1-P1' reactive center (R346 M347) acts as a "bait" for tight binding to t-PA/urokinase-type PA. In vivo, PAI-1 is encountered in complex with vitronectin, an interaction known to stabilize its activity but not to affect the second-order association rate constant (k1) between PAI-1 and t-PA. Nevertheless, by using PAI-1 reactive site variants (R346M, M347S, and R346M M347S), we show that the binding of vitronectin to the PAI-1 mutant proteins improves plasminogen activator inhibition. In the absence of vitronectin the PAI-1 R346M mutants are virtually inactive toward t-PA (k1 less than 1 x 10(3) M-1 s-1). In contrast, in the presence of vitronectin the rate of association increases about 1,000-fold (k1 of 6-8 x 10(5) M-1 s-1). This inhibition coincides with the formation of serpin-typical, sodium dodecyl sulfide-stable t-PA.PAI-1 R346M (R346M M347S) complexes. As evidenced by amino acid sequence analysis, the newly created M346-M/S347 peptide bond is susceptible to attack by t-PA, similar to the wild-type R346-M347 peptide bond, indicating that in the presence of vitronectin M346 functions as an efficient P1 residue. In addition, we show that the inhibition of t-PA and urokinase-type PA by PAI-1 mutant proteins is accelerated by the presence of the nonprotease A chains of the plasminogen activators.  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