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.
The rat pituitary contains an enzyme which will acetylate certain corticotropin (ACTH) fragments using acetyl coenzyme A (AcCoA). This acetyltransferase activity was found in all three lobes of the rat pituitary. The enzyme was largely particulate in nature. The enzyme sedimenting at 27,000 and 100,000g had specific activities 4–10 times greater than the soluble fraction. The acetyltransferase activity was dependent on substrate concentration (ACTH), was linear with time, and was inactivated at 55 °C. The enzyme would acetylate ACTH (1–24), (1–10), and (4–10), but would not use ACTH (2–10), (3–10), or (1–8) as substrates. The apparent Km values for the substrates were as follows: AcCoA, 2.2 μm ACTH (1–24), 4.2 μm; ACTH (1–10), 96 μm; and ACTH (4–10), 37 μm.  相似文献   

2.
A hydroxynitrile lyase was isolated and purified to homogeneity from seeds of Eriobotrya japonica (loquat). The final yield, of 36% with 49-fold purification, was obtained by 30-80% (NH(4))(2)SO(4) fractionation and column chromatography on DEAE-Toyopearl and Concanavalin A Sepharose 4B, which suggested the presence of a carbohydrate side chain. The purified enzyme was a monomer with a molecular mass of 72 kDa as determined by gel filtration, and 62.3 kDa as determined by SDS-gel electrophoresis. The N-terminal sequence is reported. The enzyme was a flavoprotein containing FAD as a prosthetic group, and it exhibited a K(m) of 161 microM and a k(cat)/K(m) of 348 s(-1) mM(-1) for mandelonitrile. The optimum pH and temperature were pH 5.5 and 40 degrees C respectively. The enzyme showed excellent stability with regard to pH and temperature. Metal ions were not required for its activity, while activity was significantly inhibited by CuSO(4), HgCl(2), AgNO(3), FeCl(3), beta-mercaptoethanol, iodoacetic acid, phenylmethylsulfonylfluoride, and diethylpyrocarbonate. The specificity constant (k(cat)/K(m)) of the enzyme was investigated for the first time using various aldehydes as substrates. The enzyme was active toward aromatic and aliphatic aldehydes, and showed a preference for smaller substrates over bulky one.  相似文献   

3.
Wild-type Aerobacter aerogenes 1033 is unable to utilize xylitol. A succession of mutants was isolated capable of growth on this compound (0.2%) at progressively faster rates. Whereas the ability to utilize xylitol was achieved in the first-stage mutant (X1) by constitutive production of ribitol dehydrogenase (for which xylitol is a substrate but not an inducer), the basis for enhanced utilization of xylitol in the second-stage mutant (X2) was an alteration of ribitol dehydrogenase. This enzyme was purified from the various mutants. The apparent K(m) for xylitol was 0.12 m with X2 enzyme and 0.29 m with X1 enzyme. The X2 enzyme was also less heat stable and, at 0.05 m substrate concentration, had a higher ratio of activity with xylitol compared to ribitol than did the X1 enzyme. The third mutant (X3), with an even faster growth rate on xylitol, produced a ribitol dehydrogenase indistinguishable physically or kinetically from that of X2. However, X3 produced constitutively an active transport system which accepts xylitol. The usual function of this system is apparently for the transport of d-arabitol since the latter is not only a substrate but also an inducer of the transport system in parental strains of X3. The sequence of mutations described herein illustrates how genes belonging to different metabolic systems can be mobilized to serve a new biochemical pathway.  相似文献   

4.
Cytokinin oxidase/dehydrogenase (CKO/CKX) is a flavoenzyme, which irreversibly inactivates cytokinins by severing the isoprenoid side chain from the adenine/adenosine moiety. There are several genes coding for the enzyme in maize (Zea mays). A Z. mays CKO1 cDNA was cloned in the yeast Yarrowia lipolytica to achieve heterologous protein expression. The recombinant ZmCKO1 was recovered from cultures of transformed yeasts and purified using several chromatographic steps. The enzyme was obtained as a homogeneous protein in a remarkably high-yield and its molecular and kinetic properties were characterized. The enzyme showed a molecular mass of 69 kDa, pI was 6.3. Neutral sugar content of the molecule was 22%. Absorption and fluorescence spectra were in accordance with the presence of FAD as a cofactor. Peptide mass fingerprinting using MALDI-MS correctly assigned the enzyme in MSDB protein database. The enzyme showed a relatively high degree of thermostability (T50=55 degrees C for 30 min incubation). The following pH optimum and K(m) values were determined for natural substrates (measured in the oxidase mode): pH 8.0 for isopentenyl adenine (K(m)=0.5 microM), pH 7.6 for isopentenyl adenosine (K(m)=1.9 microM), pH 7.9 for zeatin (K(m)=1.5 microM) and pH 7.3 for zeatin riboside (K(m)=2.0 microM). ZmCKO1, functioning in the oxidase mode, catalyzes the production of one molecule of H2O2 per one molecule of cytokinin substrate. This finding represents clear evidence for the existence of dual enzyme functionality (oxygen serves as a cosubstrate in the absence of better electron acceptors).  相似文献   

5.
Open reading frame 1 of Bamboo mosaic virus (BaMV), a Potexvirus in the alphavirus-like superfamily, encodes a 155-kDa replicase responsible for the formation of the 5' cap structure and replication of the viral RNA genome. The N-terminal domain of the viral replicase functions as an mRNA capping enzyme, which exhibits both GTP methyltransferase and S-adenosylmethionine (AdoMet)-dependent guanylyltransferase activities. We mutated each of the four conserved amino acids among the capping enzymes of members within alphavirus-like superfamily and a dozen of other residues to gain insight into the structure-function relationship of the viral enzyme. The mutant enzymes were purified and subsequently characterized. H68A, the mutant enzyme bearing a substitution at the conserved histidine residue, has an approximately 10-fold increase in GTP methyltransferase activity but completely loses the ability to form the covalent m(7)GMP-enzyme intermediate. High-pressure liquid chromatography analysis confirmed the production of m(7)GTP by the GTP methyltransferase activity of H68A. Furthermore, the produced m(7)GTP sustained the formation of the m(7)GMP-enzyme intermediate for the wild-type enzyme in the presence of S-adenosylhomocysteine (AdoHcy), suggesting that the previously observed AdoMet-dependent guanylation of the enzyme using GTP results from reactions of GTP methylation and subsequently guanylation of the enzyme using m(7)GTP. Mutations occurred at the other three conserved residues (D122, R125, and Y213), and H66 resulted in abolition of activities for both GTP methylation and formation of the covalent m(7)GMP-enzyme intermediate. Mutations of amino acids such as K121, C234, D310, W312, R316, K344, W406, and K409 decreased both activities by various degrees, and the extents of mutational effects follow similar trends. The affinity to AdoMet of the various BaMV capping enzymes, except H68A, was found in good correlations with not only the magnitude of GTP methyltransferase activity but also the capability of forming the m(7)GMP-enzyme intermediate. Taken together with the AdoHcy dependence of guanylation of the enzyme using m(7)GTP, a basic working mechanism, with the contents of critical roles played by the binding of AdoMet/AdoHcy, of the BaMV capping enzyme is proposed and discussed.  相似文献   

6.
Thr(93), Ser(94), Thr(140), and Ser(306) are conserved in all adenylosuccinate lyases (ASL) and are close to other amino acids previously identified by mutagenesis as being in the active site. To test their involvement in the enzyme's function, each of these amino acids was replaced by alanine. All the mutants exhibit circular dichroism spectra which are similar to that of wild-type enzyme, indicating there is no appreciable change in secondary structure. T93A exhibits 0.5% of the V(max) of wild-type ASL with a 10-fold increase in K(m) for adenylosuccinate. S94A has 65% of the V(max) of wild-type ASL with little change in K(m). T140A exhibits 0.03% of the activity of wild-type enzyme with an 11-fold increase in K(m). S306A has 0.4% of the V(max) of wild-type ASL with a sevenfold increase in K(m). Measurements of the pH-V(max) profile reveal a pK(2) value for S94A of 7.83 and S306A of 7.65, in contrast to 8.24 for the wild-type enzyme and 8.42 for T93A. Thr(93) may orient adenylosuccinate optimally for catalysis, while Ser(94) stabilizes protonated His(89), a determinant of pK(2). Thr(140) may, through hydrogen bonding, interact with Asn(270), an amino acid essential for catalysis. Ser(306) may be involved in a hydrogen bond network that ultimately stabilizes protonated His(68), which is probably the general acid in the reaction of enzyme with substrate. The results of this paper demonstrate the importance in the catalytic function of ASL of hydrogen bonds and hydrogen bonding networks involving serine and threonine.  相似文献   

7.
Sucrose-phosphate (SPS) from source leaves of soybean ( Glycine max (L.) Merr. cv. Ransom II) was purified 74-fold to a final specific activity of 1.8 U (mg protein)1. The partially purified preparation was free from phosphoglucoseisomerase (EC 5.3.1.9), pyrophosphatase (EC 3.6.1.1), phosphoenolpyruvate-phosphatase (EC 3.1.3.-), phosphofructokinase (EC 2.7.1.11), and uridine diphosphatase (EC 3.6.1.6), and was used for characterization of the kinetic and regulatory properties of the enzyme. The enzyme showed hyperbolic saturation kinetics for both fructose-6-phosphate (Km=0.57 m M ) and UDPGlucose (UDPG) (Km=4.8 m M ). The activity of SPS was inhibited by the product UDP. In vitro this inhibition could be partially overcome by the presence of Mg2+. Inorganic orthophosphate was only slightly inhibitory (35% inhibition at 25 m M phosphate). Glucose-6-phosphate (up to 20 m M ) had no effect on activity, and did not show any significant interaction with phosphate inhibition. A range of potential effectors was tested and had no effect on SPS activity: Glucose-1-phosphate, fructose-1, 6-bisphosphate, α-glycero-phosphate, dihydroxyacetone-phosphate, 3-phosphoglyceric acid, (all at 5 m M ), sucrose at 100 m M and pyrophosphate at 0.1 m M . The apparent lack of allosteric regulation of soybean SPS makes this enzyme markedly different from SPS previously characterized from spinach and maize.  相似文献   

8.
4-Coumarate:coenzyme A ligase (4CL) plays a key role in phenylpropanoid metabolism, providing precursors for a large variety of important plant secondary metabolites, such as lignin, flavonoids, and phytoalexins. Although 4CLs have been believed to be specific to plants, a gene encoding a 4CL-like enzyme which shows more than 40% identity in amino acid sequence to plant 4CLs was found in the genome of the gram-positive, filamentous bacterium Streptomyces coelicolor A3(2). The recombinant enzyme, produced in Escherichia coli with a histidine tag at its N-terminal end, showed distinct 4CL activity. The optimum pH and temperature of the reaction were pH 8.0 and 30 degrees C, respectively. The K(m) value for 4-coumarate and k(cat) were determined as 131 +/- 4 micro M and 0.202 +/- 0.007 s(-1), respectively. The K(m) value was comparable to those of plant 4CLs. The substrate specificity of this enzyme was, however, distinctly different from those of plant 4CLs. The enzyme efficiently converted cinnamate (K(m), 190 +/- 2 micro M; k(cat), 0.475 +/- 0.012 s(-1)), which is a very poor substrate for plant 4CLs. Furthermore, the enzyme showed only low activity toward caffeate and no activity toward ferulate, both of which are generally good substrates for plant 4CLs. The enzyme was therefore named ScCCL for S. coelicolor A3(2) cinnamate CoA ligase. To determine the amino acid residues providing the unique substrate specificity of ScCCL, eight ScCCL mutant enzymes having a mutation(s) at amino acid residues that probably line up along the substrate-binding pocket were generated. Mutant A294G used caffeate as a substrate more efficiently than ScCCL, and mutant A294G/A318G used ferulate, which ScCCL could not use as a substrate, suggesting that Ala(294) and Ala(318) are involved in substrate recognition. Furthermore, the catalytic activities of A294G and A294G/A318G toward cinnamate and 4-coumarate were greatly enhanced compared with those of the wild-type enzyme.  相似文献   

9.
To gain insight into the role of the strictly conserved histidine residue, H178, in the reaction mechanism of the methionyl aminopeptidase from Escherichia coli (EcMetAP-I), the H178A mutant enzyme was prepared. Metal-reconstituted H178A binds only one equivalent of Co(II) or Fe(II) tightly with affinities that are identical to the WT enzyme based on kinetic and isothermal titration calorimetry (ITC) data. Electronic absorption spectra of Co(II)-loaded H178A EcMetAP-I indicate that the active site divalent metal ion is pentacoordinate, identical to the WT enzyme. These data indicate that the metal binding site has not been affected by altering H178. The effect of altering H178 on activity is, in general, due to a decrease in k(cat). The k(cat) value for Co(II)-loaded H178A decreased 70-fold toward MGMM and 290-fold toward MP-p-NA compared to the WT enzyme, while k(cat) decreased 50-fold toward MGMM for the Fe(II)-loaded H178A enzyme and 140-fold toward MP-p-NA. The K(m) values for MGMM remained unaffected, while those for MP-p-NA increased approximately 2-fold for Co(II)- and Fe(II)-loaded H178A. The k(cat)/K(m) values for both Co(II)- and Fe(II)-loaded H178A toward both substrates ranged from approximately 50- to 580-fold reduction. The pH dependence of log K(m), log k(cat), and log(k(cat)/K(m)) of both WT and H178A EcMetAP-I were also obtained and are identical, within error, for H178A and WT EcMetAP-I. Therefore, H178A is catalytically important but is not required for catalysis. Assignment of one of the observed pK(a) values at 8.1 for WT EcMetAP-I was obtained from plots of molar absorptivity at lambda(max(640)) vs pH for both WT and H178A EcMetAP-I. Apparent pK(a) values of 8.1 and 7.6 were obtained for WT and H178A EcMetAP-I, respectively, and were assigned to the deprotonation of a metal-bound water molecule. The data reported herein provide support for the key elements of the previously proposed mechanism and suggest that a similar mechanism can apply to the enzyme with a single metal in the active site.  相似文献   

10.
Hung HC  Chien YC  Hsieh JY  Chang GG  Liu GY 《Biochemistry》2005,44(38):12737-12745
Human mitochondrial NAD(P)+-dependent malic enzyme is inhibited by ATP. The X-ray crystal structures have revealed that two ATP molecules occupy both the active and exo site of the enzyme, suggesting that ATP might act as an allosteric inhibitor of the enzyme. However, mutagenesis studies and kinetic evidences indicated that the catalytic activity of the enzyme is inhibited by ATP through a competitive inhibition mechanism in the active site and not in the exo site. Three amino acid residues, Arg165, Asn259, and Glu314, which are hydrogen-bonded with NAD+ or ATP, are chosen to characterize their possible roles on the inhibitory effect of ATP for the enzyme. Our kinetic data clearly demonstrate that Arg165 is essential for catalysis. The R165A enzyme had very low enzyme activity, and it was only slightly inhibited by ATP and not activated by fumarate. The values of K(m,NAD) and K(i,ATP) to both NAD+ and malate were elevated. Elimination of the guanidino side chain of R165 made the enzyme defective on the binding of NAD+ and ATP, and it caused the charge imbalance in the active site. These effects possibly caused the enzyme to malfunction on its catalytic power. The N259A enzyme was less inhibited by ATP but could be fully activated by fumarate at a similar extent compared with the wild-type enzyme. For the N259A enzyme, the value of K(i,ATP) to NAD+ but not to malate was elevated, indicating that the hydrogen bonding between ATP and the amide side chain of this residue is important for the binding stability of ATP. Removal of this side chain did not cause any harmful effect on the fumarate-induced activation of the enzyme. The E314A enzyme, however, was severely inhibited by ATP and only slightly activated by fumarate. The values of K(m,malate), K(m,NAD), and K(i,ATP) to both NAD+ and malate for E314A were reduced to about 2-7-folds compared with those of the wild-type enzyme. It can be concluded that mutation of Glu314 to Ala eliminated the repulsive effects between Glu314 and malate, NAD+, or ATP, and thus the binding affinities of malate, NAD+, and ATP in the active site of the enzyme were enhanced.  相似文献   

11.
A phytase (EC 3.1.3.8) with a high affinity for phytic acid was found in Aspergillus niger SK-57 and purified to homogeneity in four steps by using ion-exchange chromatography (two types), gel filtration, and chromatofocusing. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the purified enzyme gave a single stained band at a molecular mass of approximately 60 kDa. The Michaelis constant of the enzyme for phytic acid (18.7 +/- 4.6 microM) was statistically analyzed. In regard to the orthophosphate released from phytic acid, a significant difference between a low K(m) phytase from A. niger SK-57 and a high K(m) phytase from Aspergillus ficuum was recognized.  相似文献   

12.
The phosphoglucomutase gene from a wild type Fusarium oxysporum strain (F3), was homologously expressed, under the control of the constitutive promoter of gpdA of Aspergillus nidulans. The transformant produced elevated levels of phosphoglucomutase activity compared to the wild type, a fact that facilitated the subsequent purification procedure. The enzyme (FoPGM) was purified to homogeneity applying three anion exchange and one gel filtration chromatography steps. The native enzyme revealed a monomeric structure with a molecular mass of 60 kDa, while the isoelectric point was 3.5. FoPGM was active in pH ranged from 6.0 to 8.0, with an optimum using 3-(N-morpholino)propanesulfonic acid buffer at 7.0, while loss of activity was observed when phosphate buffer was used in the above mentioned pH range. The optimal temperature for activity was 45°C but the enzyme became unstable at temperatures above 40°C. FoPGM requires the presence of a divalent cation for its function with maximum activity being obtained with Co(2+). The apparent K(m) for Co(2+) was found to be 10 μM. The enzyme was also active with other divalent metal ions such as Mn(2+), Mg(2+), Ni(2+) and Ca(2+) but to a lesser extent. The following kinetic constants were determined: v(max), 0.74 μmol mg(protein)(-1)min(-1); k(cat), 44.2 min(-1); K(m)(G1P), 0.10mM; K(m)(G1,6 diP), 1.03 μM; k(cat)/K(m)(G1P), 443 mM(-1)min(-1) and k(cat)/K(m)(G1,6 diP), 42,860 mM(-1)min(-1). The enzyme was considered to follow a Ping Pong substituted enzyme or enzyme isomerization mechanism.  相似文献   

13.
A simple and direct assay method for glucose oxidase (EC 1.1.3.4) from Aspergillus niger and Penicillium amagasakiense was investigated by Fourier transform infrared spectroscopy. This enzyme catalyzed the oxidation of d-glucose at carbon 1 into d-glucono-1,5-lactone and hydrogen peroxide in phosphate buffer in deuterium oxide ((2)H(2)O). The intensity of the d-glucono-1,5-lactone band maximum at 1212 cm(-1) due to CO stretching vibration was measured as a function of time to study the kinetics of d-glucose oxidation. The extinction coefficient epsilon of d-glucono-1,5-lactone was determined to be 1.28 mM(-1)cm(-1). The initial velocity is proportional to the enzyme concentration by using glucose oxidase from both A. niger and P. amagasakiense either as cell-free extracts or as purified enzyme preparations. The kinetic constants (V(max), K(m), k(cat), and k(cat)/K(m)) determined by Lineweaver-Burk plot were 433.78+/-59.87U mg(-1) protein, 10.07+/-1.75 mM, 1095.07+/-151.19s(-1), and 108.74 s(-1)mM(-1), respectively. These data are in agreement with the results obtained by a spectrophotometric method using a linked assay based on horseradish peroxidase in aqueous media: 470.36+/-42.83U mg(-1) protein, 6.47+/-0.85 mM, 1187.77+/-108.16s(-1), and 183.58 s(-1)mM(-1) for V(max), K(m), k(cat), and k(cat)/K(m), respectively. Therefore, this spectroscopic method is highly suited to assay for glucose oxidase activity and its kinetic parameters by using either cell-free extracts or purified enzyme preparations with an additional advantage of performing a real-time measurement of glucose oxidase activity.  相似文献   

14.
A beta-1,3-galactosyl-N-acetylhexosamine phosphorylase (GalGlyNAcP) homolog gene was cloned from Vibrio vulnificus CMCP6. In synthetic reactions, the recombinant enzyme acted only with GlcNAc and GalNAc as acceptors in the presence of alpha-d-galactose-1-phosphate as a donor to form lacto-N-biose I (LNB) (Galbeta1 --> 3GlcNAc) and galacto-N-biose (GNB) (Galbeta1 --> 3GalNAc), respectively. GlcNAc was a much better acceptor than GalNAc. The enzyme also phosphorolysed LNB faster than it phosphorolysed GNB, and the k(cat)/K(m) for LNB was approximately 60 times higher than the k(cat)/K(m) for GNB. This result indicated that the enzyme was remarkably different from GalGlyNAcP from Bifidobacterium longum, which has similar activities with LNB and GNB, and GalGlyNAcP from Clostridium perfringens, which is a GNB-specific enzyme. The enzyme is the first LNB-specific enzyme that has been found and was designated lacto-N-biose I phosphorylase. The discovery of an LNB-specific GalGlyNAcP resulted in recategorization of bifidobacterial GalGlyNAcPs as galacto-N-biose/lacto-N-biose I phosphorylases.  相似文献   

15.
Phosphoenolpyruvate (PEP) carboxylase was purified over 400-fold from Plasmodium berghei. The purified enzyme was stable in 0.4 m potassium phosphate buffer (pH 7.4) containing 0.5 m glucose, 1 mm ethylenediaminetetraacetic acid (EDTA), and 1 mm MgCl(2). It had a molecular weight of 280,000 determined by sucrose density gradient centrifugation in this buffer, but it aggregated and was unstable in the presence of different salts or a more dilute solution of potassium phosphate. The K(m) for PEP was 2.6 mm and that for Mg(2+) was 1.3 mm. The K(m) for bicarbonate was 2 mm. Citrate, nucleotides, and EDTA inhibited the PEP carboxylase of P. berghei by decreasing the concentration of free magnesium ions, but acetyl-coenzyme A, fructose-1,6-diphosphate, and aspartate did not influence its activity. A chloroquine concentration of 1.8 x 10(-4)m inhibited the enzyme 50%.  相似文献   

16.
D A Toke  M L McClintick  G M Carman 《Biochemistry》1999,38(44):14606-14613
Diacylglycerol pyrophosphate (DGPP) phosphatase, encoded by the DPP1 gene, is a membrane-associated enzyme in the yeast Saccharomyces cerevisiae. The enzyme removes the beta phosphate from DGPP to form phosphatidate. The substrate and product of the DGPP phosphatase reaction play roles in lipid signaling and in cell metabolism. The deduced primary structure of the DGPP phosphatase protein contains a three-domain phosphatase sequence motif. In this work, we examined the hypothesis that the phosphatase sequence motif in the enzyme is involved in the DGPP phosphatase reaction. The amino acid residues Arg(125), His(169), and His(223) in domains 1, 2, and 3, respectively, of the phosphatase sequence motif were changed to alanine residues by site-directed mutagenesis. The mutant DPP1(R125A), DPP1(H169A), and DPP1(H223A) alleles were cloned into a yeast shuttle vector and then expressed in a dpp1Delta lpp1Delta double mutant that lacks DGPP phosphatase activity. Northern blot and immunoblot analyses showed that the mutations in the phosphatase sequence motif did not affect the expression of the enzyme. The DGPP phosphatase activities of the R125A, the H169A, and the H223A mutant enzymes were 0.05, 9, and 0.03%, respectively, of the DGPP phosphatase activity of the wild-type enzyme. Enzymes with mutations in more than one domain of the phosphatase sequence motif had no measurable DGPP phosphatase activity. The R125A and H233A mutant DGPP phosphatase enzymes had reduced V(max) and elevated K(m) values for DGPP when compared with the wild-type enzyme. The H169A mutant enzyme had reduced V(max) and K(m) values when compared with the control. The specificity constants (V(max)/K(m)()) for DGPP of the R125A mutant and H233A mutant enzymes were 4610-fold and 15 367-fold lower, respectively, when compared to the wild-type enzyme. The studies reported here indicated that the phosphatase sequence motif played an important role in the reaction catalyzed by the S. cerevisiae DGPP phosphatase.  相似文献   

17.
Steady-state and time-resolved emission spectroscopy were used to study the interaction of Escherichia coli purine nucleoside phosphorylase (PNP) with its specific inhibitors, viz. formycin B (FB), and formycin A (FA) and its N-methylated analogues, N(1)-methylformycin A (m(1)FA), N(2)-methylformycin A (m(2)FA) and N(6)-methylformycin A (m(6)FA), in the absence and presence of phosphate (P(i)). Complex formation led to marked quenching of enzyme tyrosine intrinsic fluorescence, with concomitant increases in fluorescence of FA and m(6)FA, independently of the presence of P(i). Fluorescence of m(1)FA in the complex increased only in the presence of P(i), while the weak fluorescence of FB appeared unaffected, independently of P(i). Analysis of the emission, excitation and absorption spectra of enzyme-ligand mixtures pointed to fluorescence resonance energy transfer (FRET) from protein tyrosine residue(s) to FA and m(6)FA base moieties, as a major mechanism of protein fluorescence quenching. With the non-inhibitor m(2)FA, fluorescence emission and excitation spectra were purely additive. Effects of enzyme-FA, or enzyme-m(6)FA, interactions on nucleoside excitation and emission spectra revealed shifts in tautomeric equilibria of the bound ligands. With FA, which exists predominantly as the N(1)-H tautomer in solution, the proton N(1)-H is shifted to N(2), independently of the presence of P(i). Complex formation with m(6)FA in the absence of P(i) led to a shift of the amino-imino equilibrium in favor of the imino species, and increased fluorescence at 350 nm; by contrast, in the presence of P(i), the equilibrium was shifted in favor of the amino species, accompanied by higher fluorescence at 430 nm, and a higher affinity for the enzyme, with a dissociation constant K(d)=0.5+/-0.1 microM, two orders of magnitude lower than that for m(6)FA in the absence of P(i) (K(d)=46+/-5 microM). The latter was confirmed by analysis of quenching of enzyme fluorescence according to a modified Stern-Volmer model. Fractional accessibility values (f(a)) varied from 0.31 for m(1)FA to 0.70 for FA, with negative cooperative binding of m(1)FA and FB, and non-cooperative binding of FA and m(6)FA. For all nucleoside ligands, the best model describing binding stoichiometry was one ligand per native enzyme hexamer. Fluorescence decays of PNP, FA and their mixtures were best fitted to a sum of two exponential terms, with average lifetimes () affected by their interactions. Complex formation resulted in a 2-fold increase in of FA, and a 2-fold decrease in of enzyme fluorescence. The amplitude of the long-lifetime component also increased, confirming the shift of the tautomeric equilibrium in favor of the N(2)-H species. The findings have been examined in relation to enzyme-nucleoside binding deduced from structural studies.  相似文献   

18.
A phage lytic enzyme was isolated from lysates of Bacillus stearothermophilus (NCA 1503-4R). The enzyme was purified 1,998-fold with a 27% recovery of enzyme activity. By use of polyacrylamide gel electrophoresis and sucrose gradient centrifugation the enzyme was judged free from protein contaminants. The lytic enzyme was active over a pH range of 6.0 to 7.0, with a maximum at 6.3, and it was stable between pH 7.0 and 8.0 and at 5.0 and unstable between pH 5.5 and 6.5. The temperature coefficient (Q(10)) was 2.27 between 35 and 45 C, 2.01 between 45 and 55 C, and 2.00 between 50 and 60 C. Lytic enzyme in 0.1 m sodium phosphate was not inactivated after a 1-hr exposure to temperatures below 65.5 C, whereas a 1% inactivation was observed at 70.6 C. A 2-hr exposure at 60.1, 65.5, and 70.6 C resulted in an inactivation of 1.2, 9.6, and 12.0%, respectively. A sodium phosphate concentration of at least 0.1 m was necessary for the prolonged exposure of lytic enzyme at 55 C (pH 6.3), whereas 0.005 m was required for maximal lytic activity. Lytic activity was stimulated 169, 165, and 160% by 10(-4)m Mg(++), Ca(++), and Mn(++), respectively. Lytic activity was inhibited 75% by 10(-4)m ethylenediaminetetraacetic acid (EDTA). The EDTA inhibition could be reversed by the addition of excess Mg(++), Ca(++), or Mn(++). Lytic activity was not affected by NaCl, KCl, or NH(4)Cl. Lytic activity was inhibited 100, 91, 25, 61, and 56% by 10(-4)m Hg(++), Cu(++), Zn(++), p-chloromercuribenzoate, and p-hydroxymercuribenzoate, respectively. Cysteine or 2-mercaptoethanol did not stimulate lytic activity, nor were these sulfhydryl compounds required for maintenance of enzyme activity during handling or storage. Cell walls were rapidly solubilized when incubated with lytic enzyme. Lytic action was complete after 1.5 min, with a 70% reduction in optical density (OD). Cell walls without lytic enzyme showed no reduction in OD during this period. The solubilization of N-terminal amino groups paralleled the reduction in OD and reached a level of 0.3 mumole/mg of cell wall after 4 min of incubation. Cell walls with and without lytic enzyme treatment showed a 3- and a 1.3-fold increase, respectively, in N-terminal amino groups after 3 hr of incubation. There was no release of reducing power in either the untreated cell wall suspensions or those treated with lytic enzyme. Electron micrographs of treated and untreated cell walls showed that the enzyme partially degrades the cell wall with the release of small wall fragments.  相似文献   

19.
Leucyl-transfer ribonucleic acid (tRNA) synthetase was purified 100-fold from extracts of Salmonella typhimurium. The partially purified enzyme had the following K(m) values: leucine, 1.1 x 10(-5)m; adenosine triphosphate, 6.5 x 10(-4)m; tRNA(I) (Leu), 4.1 x 10(-8)m; tRNA(II) (Leu), 4.3 x 10(-8)m; tRNA(III) (Leu), 5.3 x 10(-8)m; and tRNA(IV) (Leu), 2.9 x 10(-8)m. The tRNA(Leu) fractions were isolated from Salmonella bulk tRNA by chromatography on reversed-phase columns and benzoylated diethylaminoethyl cellulose. The enzyme had a pH optimum of 8.5 and an activation energy of 10,400 cal per mole, and was inactivated exponentially at 49.5 C with a first-order rate constant of 0.064 min(-1). Strain CV356 (leuS3 leuABCD702 ara-9 gal-205) was isolated as a mutant resistant to dl-4-azaleucine and able to grow at 27 C but not at 37 C. Extracts of strain CV356 had no leucyl-tRNA synthetase activity (charging assay) when assayed at 27 or 37 C. Temperature sensitivity and enzyme deficiency were caused by mutation in the structural gene locus specifying leucyl-tRNA synthetase. A prototrophic derivative of strain CV356 (CV357) excreted branched-chain amino acids and had high pathway-specific enzyme levels when grown at temperatures where its doubling time was near normal. At growth-restricting temperatures, both amino acid excretion and enzyme levels were further elevated. The properties of strain CV357 indicate that there is only a single leucyl-tRNA synthetase in S. typhimurium.  相似文献   

20.
The carminomycin 4-O-methyltransferase enzyme from Streptomyces peucetius was covalently immobilized on 3M Emphaze ABI-activated beads. Optimal conditions of time, temperature, pH, ionic strength, enzyme, substrate (carminomycin), and cosubstrate (S-adenosyl-L-methionine) concentrations were defined for the immobilization reaction. Protein immobilization yield ranged from 52% to 60%. Including carminomycin during immobilization had a positive effect on the activity of the immobilized enzyme but a strongly negative effect on the coupling efficiency. The immobilized enzyme retained at least 57% of its maximum activity after storage at 4 degrees C for more than 4 months. The properties of the free and immobilized enzyme were compared to determine whether immobilization could alter enzyme activity. Both soluble and bound enzyme exhibited the same pH profile with an optimum near 8.0. Immobilization caused an approximately 50% decrease in the apparent K(m) (K'(m)) for carminomycin while the K'(m) for S-adenosyl-L-methionine was approximately doubled. A 57% decrease in the V(max) value occurred upon immobilization. These changes are discussed in terms of active site modifications as a consequence of the enzyme immobilization. This system has a potential use in bioreactors for improving the conversion of carminomycin to daunorubicin. (c) 1995 John Wiley & Sons, Inc.  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