首页 | 官方网站   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.
Ma K  Weiss R  Adams MW 《Journal of bacteriology》2000,182(7):1864-1871
The fermentative hyperthermophile Pyrococcus furiosus contains an NADPH-utilizing, heterotetrameric (alphabetagammadelta), cytoplasmic hydrogenase (hydrogenase I) that catalyzes both H(2) production and the reduction of elemental sulfur to H(2)S. Herein is described the purification of a second enzyme of this type, hydrogenase II, from the same organism. Hydrogenase II has an M(r) of 320,000 +/- 20,000 and contains four different subunits with M(r)s of 52,000 (alpha), 39,000 (beta), 30,000 (gamma), and 24,000 (delta). The heterotetramer contained Ni (0.9 +/- 0.1 atom/mol), Fe (21 +/- 1.6 atoms/mol), and flavin adenine dinucleotide (FAD) (0.83 +/- 0.1 mol/mol). NADPH and NADH were equally efficient as electron donors for H(2) production with K(m) values near 70 microM and k(cat)/K(m) values near 350 min(-1) mM(-1). In contrast to hydrogenase I, hydrogenase II catalyzed the H(2)-dependent reduction of NAD (K(m), 128 microM; k(cat)/K(m), 770 min(-1) mM(-1)). Ferredoxin from P. furiosus was not an efficient electron carrier for either enzyme. Both H(2) and NADPH served as electron donors for the reduction of elemental sulfur (S(0)) and polysulfide by hydrogenase I and hydrogenase II, and both enzymes preferentially reduce polysulfide to sulfide rather than protons to H(2) using NADPH as the electron donor. At least two [4Fe-4S] and one [2Fe-2S] cluster were detected in hydrogenase II by electron paramagnetic resonance spectroscopy, but amino acid sequence analyses indicated a total of five [4Fe-4S] clusters (two in the beta subunit and three in the delta subunit) and one [2Fe-2S] cluster (in the gamma subunit), as well as two putative nucleotide-binding sites in the gamma subunit which are thought to bind FAD and NAD(P)(H). The amino acid sequences of the four subunits of hydrogenase II showed between 55 and 63% similarity to those of hydrogenase I. The two enzymes are present in the cytoplasm at approximately the same concentration. Hydrogenase II may become physiologically relevant at low S(0) concentrations since it has a higher affinity than hydrogenase I for both S(0) and polysulfide.  相似文献   

2.
The flavin-free azoreductase from Xenophilus azovorans KF46F (AzoB), which has been the very first characterized oxygen-tolerant azoreductase, was analyzed in comparison to various recently described flavin-containing azoreductases from different bacterial sources. Sequence comparisons demonstrated that the azoreductase from X. azovorans KF46F is a member of the NmrA family of proteins and demonstrates 30% sequence identity with a NADPH-dependent quinone oxidoreductase from Escherichia coli (encoded by ytfG). In contrast, it was found that the flavin-containing azoreductases from E. coli OY1-2 (AZR), Bacillus sp. OY1-2 (AZR) and related azoreductases all belong to the FMN_red superfamily of enzymes. The substrate specificity of AzoB was reanalyzed in respect to the recently characterized flavin-containing azoreductases, and it was found that purified AzoB converted in addition to different ortho-hydroxy azo compounds [such as Orange II = 1-(4′-sulfophenylazo)-2-naphthol] also the simple non-hydroxylated non-sulfonated azo dye Methyl Red (4′-dimethylaminoazobenzene-2-carboxylic acid), but no indications for the conversion of quinones were obtained. Significant differences were observed in the substrate specificities between AzoB and the flavin-containing azoreductases. The kinetic analysis of the turn-over of Orange II by AzoB suggested an ordered bireactant reaction mechanism which was different from the ping-pong mechanism suggested for the flavin-containing azoreductases.  相似文献   

3.
Soluble and cell wall bound gamma-glutamyltransferases (GGTs) were purified from radish (Raphanus sativus L.) cotyledons. Soluble GGTs (GGT I and II) had the same M(r) of 63,000, and were composed of a heavy subunit (M(r), 42,000) and a light one (M(r), 21,000). The properties of GGT I and II were similar. Bound GGTs (GGT A and B) were purified to homogeneity from the pellet after the extraction of soluble GGTs. GGT A and B were monomeric proteins with an M(r) of 61,000. The properties of GGT A and B were similar. Thus, bound GGTs were distinguished from soluble GGTs. The optimal pHs of soluble and bound GGTs were about 7.5. Both soluble and bound GGTs utilized glutathione, gamma-L-glutamyl-p-nitroanilide, oxidized glutathione and the conjugate of glutathione with monobromobimane as substrates, and were inhibited by acivicin, but soluble GGTs were also distinguished from bound GGTs with regard to these properties.  相似文献   

4.
Distinct metal-binding configurations in metallothionein   总被引:9,自引:0,他引:9  
In a study of the binding stoichiometry of various metals to rat liver metallothionein, the protein appears to coordinate metals in 2 distinct configurations. Ions of at least 18 different metals were shown to associate with the protein suggesting that there is little specificity in binding. Most metals exhibited saturation binding at 7 mol eq forming M7-metallothionein. These included Bi(III), Cd(II), Co(II), Hg(II), In(III), Ni(II), Pb(II), Sb(III), and Zn(II). Others metals including Os(III), Pd(II), Pt(IV), Re(V), Rh(III), and Tl(III) give a positive indication of binding, but stoichiometries were unclear. Ag(I) and Cu(I) bound in clusters as M12-metallothionein. This binding stoichiometry was determined in 3 ways: (a) by determining the equivalence point in Cu- and Ag-titrated samples where resistance to proteolysis is maximal; (b) by determining the point where Zn ions are completely displaced from Zn7-metallothionein; and (c) by direct binding studies. Ag-reconstituted protein, recovered from gel filtration, had an average Ag content of 11.5 g atoms/mol of protein. A similar stoichiometry for the Cu-protein resulted from displacement of Zn from Zn7-metallothionein by Cu(I). The M12-protein was converted to the M7-protein by displacement of Ag(I) or Cu(I) with 7 mol eq of Hg(II). Whereas the distribution of metals in the 2 domains of M7-metallothionein is M4 alpha and M3 beta, the arrangement in the M12-molecule is probably M6 alpha and M6 beta. We propose that metallothionein ligates Ag(I) and Cu(I) in a trigonal geometry by bridging thiolates. This is in contradistinction to a tetrahedral binding geometry in the M7-protein. Distinct binding configurations may result in different tertiary structures for M7- and M12-proteins which may relate to metabolic specificity of Zn-metallothionein and Cu-metallothionein, respectively.  相似文献   

5.
Two distinct DNA ligases from Drosophila melanogaster embryos   总被引:5,自引:0,他引:5  
M Takahashi  M Senshu 《FEBS letters》1987,213(2):345-352
Embryos of Drosophila melanogaster contain two distinct DNA ligases (DNA ligase I and II). DNA ligase I was eluted at 0.2 M KCl and DNA ligase II at 0.6 M KCl on phosphocellulose column chromatography. The former was rich in early developing embryos and its activity decreased during embryonic development. The latter was found constantly throughout the developing stages of embryos. DNA ligase I existed in a cytoplasmic fraction and DNA ligase II is concentrated in nuclei. Both enzymes ligate 5'-phosphoryl and 3'-hydroxyl groups in oligo(dT) in the presence of poly(dA). DNA ligase II is also able to join oligo(dT)(poly(rA). Both enzymes require ATP and Mg2+ for activity. The Km for ATP is 2.7 X 10(-6) M for DNA ligase I, and 3.0 X 10(-5) M for DNA ligase II. DNA ligase I requires dithiothreitol and polyvinyl alcohol, but DNA ligase II does not. Both enzymes are inhibited in the presence of N-ethylmaleimide. DNA ligase I is active at a low salt concentration (0-30 mM KCl), but DNA ligase II is active at high salt concentrations (50-100 mM). DNA ligase I is more labile than DNA ligase II. The molecular masses of DNA ligase-AMP adducts were determined as 86 and 75 kDa for DNA ligase I, and as 70 (major protein) and 90 kDa (minor protein) for DNA ligase II under denaturing conditions. A sedimentation coefficient of 4.2 S was observed for DNA ligase II. Consequently, Drosophila DNA ligase I and II are quite similar to mammalian DNA ligase I and II. Drosophila DNA ligase I and a DNA ligase by B.A. Rabin et al. [(1986) J. Biol. Chem. 261, 10637-10645] seem to be the same enzyme.  相似文献   

6.
The gene coding for an aerobic azoreductase was cloned from Pigmentiphaga kullae K24, which is able to grow with the carboxylated azo compound 1-(4'-carboxyphenylazo)-4-naphthol (carboxy-Orange I) as sole source of carbon and energy. The gene encoded a protein with a molecular weight of 20,557 Da, with a conserved putative NAD(P)H-binding site in the amino-terminal region. The deduced amino acid sequence showed no further significant sequence homologies to previously studied aerobic azoreductases. The azoreductase was heterologously expressed in Escherichia coli and shown to convert the sulfonated azo dye Orange I and furthermore Magneson II [4-(4-nitrophenylazo)-1-naphthol].  相似文献   

7.
The purification and kinetic characteristics of two phosphofructokinases are described. Aerobic cultures of Escherichia coli exhibit two types of phosphofructokinase. Both types are dimers of mol. wt 150,000 (subunit mol. wt 73,000), whereas the anaerobic culture of E. coli revealed only one type, which is a tetramer of mol. wt 350,000 (subunit mol. wt 90,000). Type 1 of the aerobic enzyme, representing approximately 70% of the total enzyme activity, is ATP-insensitive, whereas type II and the anaerobic enzyme are ATP-sensitive. The addition of AMP stimulates the tetramer, relieving ATP inhibition, and also the type II dimer, which is, however, inhibited at concentrations higher than 0.5 mM AMP. No effect was observed on the type I dimer of the aerobic preparation. ADP stimulates the tetramer and inhibits type I more strongly than type II of the aerobic dimer. The kinetic characteristics together with the effect of metabolites on these phosphofructokinase types are described and discussed in the light of their importance for the regulatory mechanism of the Pasteur effect.  相似文献   

8.
Two kinds of phospholipids in normal rat uterus were found to inhibit the aggregation of washed rabbit platelets induced by 1-alkyl-2-acetyl-sn-glycero-3-phosphocholine (alkylacetyl-GPC) and were named Inhibitor I and Inhibitor II and identified by mass spectrometry. Inhibitor I was a mixture of 1-acyl (16:0, 18:0, 18:1, 18:2, and 20:4)-2-lyso-sn-glycero-3-phosphocholine (acyllyso-GPC) and 1-alkyl (16:0, 18:0, and 18:1)-2-lyso-sn-glycero-3-phosphocholine (alkyllyso-GPC). 16:0 acyllyso-GPC was the most inhibitory, followed by 18:1, 18:2, 20:4, and 18:0 acyllyso-GPCs and 16:0 alkyllyso-GPC. Their IC50 values were in the range of 1-4 X 10(-5) M against the platelet aggregation induced by 1 X 10(-10) M 16:0 alkylacetyl-GPC, indicating that they were about 100 times weaker inhibitors than CV-3988. Inhibitor II was a mixture of N-acyl sphing-4-enyl phosphocholine (18:1/18:0, 18:1/20:0, 18:1/24:0, and 18:1/24:2). The most inhibitory of these components were 18:1/20:0 and 18:1/24:0, followed by 18:1/24:2 and 18:1/18:0, and their IC50 values were in the range of 4-5 X 10(-5) M against platelet aggregation induced by the alkylacetyl-GPC. Quantitatively, about 10(5) times higher concentrations of these inhibitors should be necessary to inhibit platelet aggregation induced by 1 X 10(-10) M 16:0 alkylacetyl-GPC. In fact, the contents of Inhibitors I and II, respectively, were approximately 10(5) times (4.7 X 10(-2) and 7.1 X 10(-2) mol/mol lipid-phosphorus of the original uterine phospholipids) than that of 16:0 alkylacetyl-GPC (1.4 X 10(-6) mol/mol lipid-phosphorus). The role of alkylacetyl-GPC in normal rat uterus is uncertain, but it coexists in situ with two kinds of endogenous inhibitors, choline containing lysoglycerophospholipids and sphingophospholipids.  相似文献   

9.
Specific binding of corticosteroids in cultured CCT cells was measured as a function of time, of temperature, of pH, and of concentration. Scatchard analysis revealed the existence of two species of binding sites for both, 3H-aldosterone type I A: KD = 2.3.10(-9) M, N = 33.10(-17) mol/10(4) cells; type I B: KD = 51.10(-9) M, N = 55.10(-17) mol/10(4) cells) and dexamethasone (type II A: KD = 4.7.10(-9) M, N = 2.3.10(17) mol/10(4) cells; type II B: KD = 22.10(-9) M, N = 6.5.10(-17) mol/10(4) cells). The data demonstrate that CCT cells in primary monolayer culture express corticosteroid binding sites similar to cells of the CCT in vivo.  相似文献   

10.
Two plant-type ferredoxins were isolated and purified from a blue-green alga, Nostoc verrucosum. They were separable by chromatography on a DEAE-cellulose column. The slow-moving band was designated ferredoxin I (Fd I) and the fast-moving band was ferredoxin II (Fd II). The ratio of the yield of ferredoxins I and II was about 1 : 0.84. Both ferredoxins had absorption spectra similar to those of plant-type ferredoxins. Two atoms of non-heme iron and two of labile sulfur were found per mol of both ferredoxin I and ferredoxin II. Their molecular weights were identical and estimated to be about 18 000 by a gel filtration method. The biochemical activities of these Nostoc ferredoxins were studied: the NADP photoreduction activity on one hand and the NADP-cytochrome c reductase activity on the other.  相似文献   

11.
Photosynthetic fructose-1,6-diphosphatase (FDPase) fractions I and II, earlier purified from spinach leaves, show a similar amino acid composition, with the exception of a higher glutamic acid content in the latter. In both fractions glutamic and aspartic acids are the main amino acids. pH activity profiles of fractions I and II are similar, with optima at 8·65–8·70, both showing a high specificity for fructose- 1,6-diphosphate. These two fractions are Mg2+-dependent for activity, with an Optimum Mg2+ concentration of 10 mM in standard conditions, which shifts to 5 mM when the MG2+/EDTA ratio is increased to 10; Mn2+ and Co2+ are slightly active. EDTA enhances FDPase activity slightly, with an optimum at 0·4–0·8 mM. Cysteine has no activating effect, and acts as an inhibitor above 10 mM. Both I and II have an optimum substrate concentration of 4 mM, and the substrate inhibits at concns above this value. Kinetic velocity curves are sigmoidal, with the concave zone located in the range of physiological substrate concns. (Hill coefficient 1·75 for both). This suggests a strong regulatory role of fructose-1,6-diphosphate. Km values are 1·4 × 10−3 M (fraction I) and 1·1 × 10−3 M (fraction II). The highest activity rate occurs at 60°, in accordance with the high thermostability of both fractions; the activation energies are 14·3 kcal/mol (fraction I) and 13·0 kcal/mol (fraction II).  相似文献   

12.
A protease-producing, crude oil degrading marine isolate was identified as Nocardiopsis sp. on the basis of the morphology, cell wall composition, mycolic acid analysis and DNA base composition. The Nocardiopsis produces two extracellular proteases, both of which are alkaline serine endopeptidases. Protease I was purified to homogeneity by chromatography on CM-Sephadex at pH 5.0 and pH 9.0. Protease II was purified using DEAE-cellulose, Sephadex G-50, phenyl-Sepharose and hydroxyapatite chromatography. Protease I and II had almost similar M(r) of 21 kDa (Protease I) and 23 kDa (Protease II), pI of 8.3 and 7.0 respectively with pH and temperature optima for activity between 10.0 and 11.0 and about 60 degrees C. Specific activities were 152 and 14 U/mg respectively on casein. However, Protease I was antigenically unrelated to Protease II. Both proteases were endopeptidases and required extended substrate binding for catalysis. Both proteases had collagenolytic and fibrinolytic activity but only Protease I had elastinolytic activity. The proteases were chymotrypsin-like with respect to their amino acid compositions and N-terminal sequences.  相似文献   

13.
Two plant-type ferredoxins were isolated and purified from a blue-green alga, Nostoc verrucosum. They were separable by chromatography on a DEAE-cellulose column. The slow-moving band was designated ferredoxin I (Fd I) and the fast-moving band was ferredoxin II (Fd II). The ratio of the yield of ferredoxins I and II was about 1:0.84. Both ferredoxins had absorption spectra similar to those of plant-type ferredoxins. Two atoms of non-heme iron and two of labile sulfur were found per mol of both ferredoxin I and ferredoxin II. Their molecular weights were identical and estimated to be about 18 000 by a gel filtration method. The biochemical activities of these Nostoc ferredoxins were studied: the NADP photoreduction activity on one hand and the NADP-cytochrome c reductase activity on the other.  相似文献   

14.
Trehalases from a thermophilic fungus Thermomyces lanuginosus (M(r) 145 kDa) and a mesophilic fungus Neurospora crassa (M(r) 437 kDa) were purified to compare their thermal characteristics and kinetic constants. Both trehalases were maximally active at 50 degrees C, had an acidic pH optimum and were glycoproteins (20% and 43%, w/w, carbohydrate content for T. lanuginosus and N. crassa, respectively). At their temperature optimum, their K(m) was similar (0.57 and 0.52 mM trehalose, for T. lanuginosus and N. crassa, respectively) but the V(max) of N. crassa enzyme was nine times higher than of T. lanuginosus enzyme. The catalytic efficiency, k(cat)/K(m), for N. crassa trehalase was one order of magnitude higher (6.2 x 10(6) M(-1) s(-1)) than of T. lanuginosus trehalase (4 x 10(5) M(-1) s(-1)). At their T(opt) (50 degrees C), trehalase from both sources exhibited similar thermostability (t(1/2)6 h). The energy of activation, E(a), for T. lanuginosus trehalase was 15.12 kcal mol(-1) and for N. crassa trehalase it was 9.62 kcal mol(-1). The activation energy for thermal inactivation for the N. crassa enzyme (92 kcal mol(-1)) was two-fold higher than for the T. lanuginosus enzyme (46 kcal mol(-1)). The present study shows that the trehalase of N. crassa is not only more stable but also a better catalyst than the T. lanuginosus enzyme.  相似文献   

15.
Amaranth seed proteins have a better balance of essential amino acids than cereals and legumes. In addition, the tryptic hydrolysis of amaranth proteins generates, among other peptides, angiotensin converting enzyme (ACE) inhibitory (ACEi) peptides. ACE converts angiotensin I (Ang I) into Ang II, but is also responsible for the degradation of bradykinin (BK). In contrast to Ang II, BK stimulates vasodilation modulated through endothelial nitric oxide (NO) production. The aim of the present study was to characterize the ACEi activity of amaranth trypsin-digested glutelins (TDGs) and their ability to induce endothelial NO production. An IC50 value of 200 μg ml?1 was measured for TDG inhibition of ACE. TDGs stimulated endothelial NO production in coronary endothelial cells (CEC) by 52% compared to control. The effects of TDGs were comparable to those of BK and Captopril, both used as positive controls of NO production. Consistent with these effects, TDGs induced, in a dose-dependent manner, endothelial NO-dependent vasodilation in isolated rat aortic rings. These results suggest that TDGs induce endothelial NO production and consequent vasodilation through their ACEi activity. Amaranth TDGs have a high potential as a nutraceutical food in prevention of cardiovascular diseases. Further molecular, cellular and physiological studies are currently under way and the results may contribute to a better understanding and control of cardiovascular disorders.  相似文献   

16.
Both the triple-helical and denatured forms of nonfibrillar bovine dermal type I collagen were tested as substrates for the catalytic subunit of cAMP-dependent protein kinase in an in vitro reaction. Native, triple-helical collagen was not phosphorylated, but collagen that had been thermally denatured into individual alpha chains was a substrate for the protein kinase. Catalytic subunit of cAMP-dependent protein kinase phosphorylated denatured collagen to between 3 to 4 mol of phosphate/mol of (alpha 1(I)2 alpha 2(I). Pepsin-solubilized and intact collagens were phosphorylated similarly, as long as each was in a nonhelical conformation. The first 2 mol of phosphate incorporated into type I collagen by the protein kinase were present in the alpha 2(I) chain. The alpha 1(I) chain was only phosphorylated during long incubations in which the stoichiometry exceeded 2 mol of phosphate/mol of (alpha 1(I)2 alpha 2(I). Phosphoserine was the only phosphoamino acid identified in collagen that had been phosphorylated to any degree by the protein kinase. The 2 mol of phosphate incorporated into the alpha 2(I) chain were localized to the alpha 2(I)CB4 cyanogen bromide fragment. The catalytic subunit of cAMP-dependent protein kinase phosphorylated denatured pepsin-solubilized collagen with a Km of 8 microM and a Vmax of approximately 0.1 mumol/min/mg of enzyme. Denatured, but not triple-helical, type I collagen was also phosphorylated by cGMP-dependent protein kinase, although it was a poorer substrate for this enzyme than for the cAMP-dependent protein kinase. Collagen was not a substrate for phospholipid-sensitive Ca2+-dependent protein kinase. These results suggest the potential for nascent alpha chains of type I collagen to be susceptible to phosphorylation by cAMP-dependent protein kinase in vivo prior to triple-helix formation. Such a phosphorylation of collagen could be relevant to the action of cAMP to increase the intracellular degradation of newly synthesized collagen.  相似文献   

17.
The [Ru(II)(Hedta)NO(+)] complex is a diamagnetic species crystallizing in a distorted octahedral geometry, with the Ru-N(O) length 1.756(4) A and the RuNO angle 172.3(4) degrees . The complex contains one protonated carboxylate (pK(a)=2.7+/-0.1). The [Ru(II)(Hedta)NO(+)] complex undergoes a nitrosyl-centered one-electron reduction (chemical or electrochemical), with E(NO+/NO)=-0.31 V vs SCE (I=0.2 M, pH 1), yielding [Ru(II)(Hedta)NO](-), which aquates slowly: k(-NO)=2.1+/-0.4x10(-3) s(-1) (pH 1.0, I=0.2 M, CF(3)COOH/NaCF(3)COO, 25 degrees C). At pHs>12, the predominant species, [Ru(II)(edta)NO](-), reacts according to [Ru(II)(edta)NO](-)+2OH(-)-->[Ru(II)(edta)NO(2)](3-), with K(eq)=1.0+/-0.4 x 10(3) M(-2) (I=1.0 M, NaCl; T=25.0+/-0.1 degrees C). The rate-law is first order in each of the reactants for most reaction conditions, with k(OH(-))=4.35+/-0.02 M(-1)s(-1) (25.0 degrees C), assignable mechanistically to the elementary step comprising the attack of one OH(-) on [Ru(II)(edta)NO](-), with subsequent fast deprotonation of the [Ru(II)(edta)NO(2)H](2-) intermediate. The activation parameters were DeltaH(#)=60+/-1 kJ/mol, DeltaS(#)=-31+/-3 J/Kmol, consistent with a nucleophilic addition process between likely charged ions. In the toxicity up-and-down tests performed with Swiss mice, no death was observed in all the doses administered (3-9.08 x 10(-5) mol/kg). The biodistribution tests performed with Wistar male rats showed metal in the liver, kidney, urine and plasma. Eight hours after the injection no metal was detected in the samples. The vasodilator effect of [Ru(II)(edta)NO](-) was studied in aortic rings without endothelium, and was compared with sodium nitroprusside (SNP). The times of maximal effects of [Ru(II)(edta)NO](-) and SNP were 2 h and 12 min, respectively, suggesting that [Ru(II)(edta)NO](-) releases NO slowly to the medium in comparison with SNP.  相似文献   

18.
Pentamannose 6-phosphate/trilysine substituted aprotinin (PMP-lys-aprotinin) and insulin like growth factor II (IGF II) were used as affinity ligands for the mannose 6-phosphate (M6P) and IGF II binding sites of the M6P/IGF II receptor. Both ligands were cross linked to intact receptor and tryptic fragments of the receptor. The pattern of receptor fragments with M6P and IGF II binding sites differed indicating that the two binding sites are located on different segments of the receptor. The receptor was incubated with [125I]IGF II and pentamannose 6-phosphate substituted bovine serum albumin (PMP-BSA). From these mixtures [125I]IGF II receptor complexes could be precipitated with antibodies against the PMP-BSA indicating that the M6P/IGF II receptor can bind simultaneously IGF II and M6P-containing ligands.  相似文献   

19.
Domain II (residues 189-338, M(r) = 16 222) of glutamate dehydrogenase from the hyperthermophilic bacterium Thermotoga maritima was used as a model system to study reversible unfolding thermodynamics of this hyperthermostable enzyme. The protein was produced in large quantities in E.COLI: using a T7 expression system. It was shown that the recombinant domain is monomeric in solution and that it comprises secondary structural elements similar to those observed in the crystal structure of the hexameric enzyme.The recombinant domain is thermostable and undergoes reversible and cooperative thermal unfolding in the pH range 5.90-8.00 with melting temperatures between 75.1 and 68.0 degrees C. Thermal unfolding of the protein was studied using differential scanning calorimetry and circular dichroism spectroscopy. Both methods yielded comparable values. The analysis revealed an unfolding enthalpy at 70 degrees C of 70.2 +/- 4.0 kcal/mol and a DeltaC(p) value of 1.4 +/- 0.3 kcal/mol K. Chemical unfolding of the recombinant domain resulted in m values of 3.36 +/- 0.10 kcal/mol M for unfolding in guanidinium chloride and 1.46 +/- 0.04 kcal/mol M in urea. The thermodynamic parameters for thermal and chemical unfolding equilibria indicate that domain II from T.MARITIMA: glutamate dehydrogenase is a thermostable protein with a DeltaG(max) of 3.70 kcal/mol. However, the thermal and chemical stabilities of the domain are lower than those of the hexameric protein, indicating that interdomain interactions must play a significant role in the stabilization of T. MARITIMA: domain II glutamate dehydrogenase.  相似文献   

20.
Two monofunctional NiFeS carbon monoxide (CO) dehydrogenases, designated CODH I and CODH II, were purified to homogeneity from the anaerobic CO-utilizing eubacterium Carboxydothermus hydrogenoformans. Both enzymes differ in their subunit molecular masses, N-terminal sequences, peptide maps, and immunological reactivities. Immunogold labeling of ultrathin sections revealed both CODHs in association with the inner aspect of the cytoplasmic membrane. Both enzymes catalyze the reaction CO + H(2)O --> CO(2) + 2 e(-) + 2 H(+). Oxidized viologen dyes are effective electron acceptors. The specific enzyme activities were 15,756 (CODH I) and 13,828 (CODH II) micromol of CO oxidized min(-1) mg(-1) of protein (methyl viologen, pH 8.0, 70 degrees C). The two enzymes oxidize CO very efficiently, as indicated by k(cat)/K(m) values at 70 degrees C of 1.3. 10(9) M(-1) CO s(-1) (CODH I) and 1.7. 10(9) M(-1) CO s(-1) (CODH II). The apparent K(m) values at pH 8.0 and 70 degrees C are 30 and 18 microM CO for CODH I and CODH II, respectively. Acetyl coenzyme A synthase activity is not associated with the enzymes. CODH I (125 kDa, 62.5-kDa subunit) and CODH II (129 kDa, 64.5-kDa subunit) are homodimers containing 1.3 to 1.4 and 1.7 atoms of Ni, 20 to 22 and 20 to 24 atoms of Fe, and 22 and 19 atoms of acid-labile sulfur, respectively. Electron paramagnetic resonance (EPR) spectroscopy revealed signals indicative of [4Fe-4S] clusters. Ni was EPR silent under any conditions tested. It is proposed that CODH I is involved in energy generation and that CODH II serves in anabolic functions.  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-23

京公网安备 11010802026262号