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 20 毫秒
1.
Lee DW  Hong YH  Choe EA  Lee SJ  Kim SB  Lee HS  Oh JW  Shin HH  Pyun YR 《FEBS letters》2005,579(5):1261-1266
To gain insight into the structural stability of homologous homo-tetrameric l-arabinose isomerases (AI), we have examined the isothermal guanidine hydrochloride (GdnHCl)-induced unfolding of AIs from mesophilic Bacillus halodurans (BHAI), thermophilic Geobacillus stearothermophilus (GSAI), and hyperthermophilic Thermotoga maritima (TMAI) using circular dichroism spectroscopy. The GdnHCl-induced unfolding of the AIs can be well described by a two-state reaction between native tetramers and unfolded monomers, which directly confirms the validity of the linear extrapolation method to obtain the intrinsic stabilities of these proteins. The resulting unfolding free energy (DeltaGU) values of the AIs as a function of temperature were fit to the Gibbs-Helmholtz equation to determine their thermodynamic parameters based on a two-state mechanism. Compared with the stability curves of BHAI in the presence and absence of Mn2+, those of holo GSAI and TMAI were more broadened than those of the apo enzymes at all temperatures, indicating increased melting temperatures (Tm) due to decreased heat capacity (DeltaGp). Moreover, the extent of difference in DeltaCp between the apo and holo thermophilic AIs is larger than that of BHAI. From these studies, we suggest that the metal dependence of the thermophilic AIs, resulting in the reduced DeltaCp, may play a significant role in structural stability compared to their mesophilic analogues, and that the extent of metal dependence of AI stability seems to be highly correlated to oligomerization.  相似文献   

2.
The araA gene, encoding l-arabinose isomerase (AI), from the thermophilic bacterium Geobacillus thermodenitrificans was cloned and expressed in Escherichia coli. Recombinant AI was isolated with a final purity of about 97% and a final specific activity of 2.10 U/mg. The molecular mass of the purified AI was estimated to be about 230 kDa to be a tetramer composed of identical subunits. The AI exhibited maximum activity at 70 degrees C and pH 8.5 in the presence of Mn2+. The enzyme was stable at temperatures below 60 degrees C and within the pH range 7.5-8.0. d-Galactose and l-arabinose as substrate were isomerized with high activities. Ribitol was the strongest competitive inhibitor of AI with a Ki of 5.5mM. The apparent Km and Vmax for L-arabinose were 142 mM and 86 U/mg, respectively, whereas those for d-galactose were 408 mM and 6.9 U/mg, respectively. The catalytic efficiency (kcat/Km) was 48 mM(-1)min(-1) for L-arabinose and 0.5mM(-1)min(-1) for D-galactose. Mn2+ was a competitive activator and increased the thermal stability of the AI. The D-tagatose yield produced by AI from d-galactose was 46% without the addition of Mn2+ and 48% with Mn2+ after 300 min at 65 degrees C.  相似文献   

3.
Melting behaviour of D-sucrose, D-glucose and D-fructose   总被引:1,自引:0,他引:1  
The melting behaviour of d-sucrose, d-glucose and d-fructose was studied. The melting peaks were determined with DSC and the start of decomposition was studied with TG at different rates of heating. In addition, melting points were determined with a melting point apparatus. The samples were identified as d-sucrose, alpha-d-glucopyranose and beta-d-fructopyranose by powder diffraction measurements. There were differences in melting between the different samples of the same sugar and the rate of heating had a remarkable effect on the melting behaviour. For example, T(o), DeltaH(f) and T(i) (initial temperature of decomposition) at a 1 degrees Cmin(-1) rate of heating were 184.5 degrees C, 126.6Jg(-1) and 171.3 degrees C for d-sucrose, 146.5 degrees C, 185.4Jg(-1) and 152.0 degrees C for d-glucose and 112.7 degrees C, 154.1Jg(-1) and 113.9 degrees C for d-fructose. The same parameters at 10 degrees Cmin(-1) rate of heating were 188.9 degrees C, 134.4Jg(-1) and 189.2 degrees C for d-sucrose, 155.2 degrees C, 194.3Jg(-1) and 170.3 degrees C for d-glucose and 125.7 degrees C, 176.7Jg(-1) and 136.8 degrees C d-fructose. At slow rates of heating, there were substantial differences between the different samples of the same sugar. The melting point determination is a sensitive method for the characterization of crystal quality but it cannot be used alone for the identification of sugar samples in all cases. Therefore, the melting point method should be validated for different sugars.  相似文献   

4.
The araA gene encoding L-arabinose isomerase (AI) from the hyperthermophilic bacterium Thermotoga maritima was cloned and overexpressed in Escherichia coli as a fusion protein containing a C-terminal hexahistidine sequence. This gene encodes a 497-amino-acid protein with a calculated molecular weight of 56,658. The recombinant enzyme was purified to homogeneity by heat precipitation followed by Ni(2+) affinity chromatography. The native enzyme was estimated by gel filtration chromatography to be a homotetramer with a molecular mass of 232 kDa. The purified recombinant enzyme had an isoelectric point of 5.7 and exhibited maximal activity at 90 degrees C and pH 7.5 under the assay conditions used. Its apparent K(m) values for L-arabinose and D-galactose were 31 and 60 mM, respectively; the apparent V(max) values (at 90 degrees C) were 41.3 U/mg (L-arabinose) and 8.9 U/mg (D-galactose), and the catalytic efficiencies (k(cat)/K(m)) of the enzyme were 74.8 mM(-1).min(-1) (L-arabinose) and 8.5 mM(-1).min(-1) (D-galactose). Although the T. maritima AI exhibited high levels of amino acid sequence similarity (>70%) to other heat-labile mesophilic AIs, it had greater thermostability and higher catalytic efficiency than its mesophilic counterparts at elevated temperatures. In addition, it was more thermostable in the presence of Mn(2+) and/or Co(2+) than in the absence of these ions. The enzyme carried out the isomerization of D-galactose to D-tagatose with a conversion yield of 56% for 6 h at 80 degrees C.  相似文献   

5.
The Helicobacter pylori NCTC 11637 alanine racemase gene, alr1, was cloned based on a putative alanine racemase gene, alr, of H. pylori 26695. The protein, Alr1, was purified to homogeneity from Escherichia coli MB2795 cells harboring the alr1 gene. The protein exclusively catalyzes the conversion of l-alanine to the d-isomer with K(m) and V(max) values of 100 mM and 909 mumol min(-1) mg(-1), respectively. The values are 16-fold higher than those for the reaction in the reverse direction. The molecular weight of Alr1 is 42,000 by SDS-PAGE, and 68,000 by gel-filtration analysis. The optimal pH and temperature are pH 8.3 and 37 degrees C, respectively, in good accordance with the characteristics shown by the alanine racemase purified from H. pylori NCTC 11637 cells. Pyridoxal 5'-phosphate was suggested to be the cofactor. The physiological function of Alr1 is discussed regarding energy production in the microbial cells.  相似文献   

6.
The araA gene encoding L-arabinose isomerase from Bacillus stearothermophilus US100 strain was cloned, sequenced and over-expressed in E. coli. This gene encodes a 496-amino acid protein with a calculated molecular weight of 56.161 kDa. Its amino acid sequence displays the highest identity with L-AI from Thermus sp. IM6501 (98%) and that of Geobacillus stearothermophilus T6 (97%). According to SDS-PAGE analysis, under reducing and non-reducing conditions, the recombinant enzyme has an apparent molecular weight of nearly 225 kDa, composed of four identical 56-kDa subunits. The L-AI US100 was optimally active at pH 7.5 and 80 degrees C. It was distinguishable by its behavior towards divalent ions. Indeed, the L-AI US100 activity and thermostability were totally independent for metallic ions until 65 degrees C. At temperatures above 65 degrees C, the enzyme was also independent for metallic ions for its activity but its thermostability was obviously improved in presence of only 0.2 mM Co2+ and 1 mM Mn2+. The V(max) values were calculated to be 41.3 U/mg for L-arabinose and 8.9 U/mg for D-galactose. Their catalytic efficiencies (k(cat)/K(m)) for l-arabinose and D-galactose were, respectively, 71.4 and 8.46 mM(-1) min(-1). L-AI US100 converted the d-galactose into D-tagatose with a high conversion rate of 48% after 7 h at 70 degrees C.  相似文献   

7.
Recombinant LytB protein from the thermophilic eubacterium Aquifex aeolicus produced in Escherichia coli was purified to apparent homogeneity. The purified LytB protein catalyzed the reduction of (E)-4-hydroxy-3-methyl-but-2-enyl diphosphate (HMBPP) in a defined in vitro system. The reaction products were identified as isopentenyl diphosphate and dimethylallyl diphosphate. A spectrophotometric assay was established to determine the steady-state kinetic parameters of LytB protein. The maximal specific activity of 6.6+/-0.3 micromol x min(-1) x mg(-1) protein was determined at pH 7.5 and 60 degrees C. The k(cat) value of the LytB protein was 3.7+/-0.2 s(-1) and the K(m) value for HMBPP was 590+/-60 microM.  相似文献   

8.
An efficient and practical route for the large-scale synthesis of 2-deoxy-L-erythro-pentose (2-deoxy-L-ribose) starting from L-arabinose was developed using Barton-type free-radical deoxygenation reaction as a key step. The radical precursor, a phenoxythiocarbonyl ester, was prepared in situ, and the most efficient deoxygenation was achieved by slow addition of tributyltin hydride to the reaction mixture.  相似文献   

9.
l-Xylulose was used as a raw material for the production of l-xylose with a recombinantly produced Escherichia colil-fucose isomerase as the catalyst. The enzyme had a very alkaline pH optimum (over 10.5) and displayed Michaelis-Menten kinetics for l-xylulose with a Km of 41 mM and a Vmax of 0.23 μmol/(mg min). The half-lives determined for the enzyme at 35 °C and at 45 °C were 6 h 50 min and 1 h 31 min, respectively. The reaction equilibrium between l-xylulose and l-xylose was 15:85 at 35 °C and thus favored the formation of l-xylose. Contrary to the l-rhamnose isomerase catalyzed reaction described previously [14]l-lyxose was not detected in the reaction mixture with l-fucose isomerase. Although xylitol acted as an inhibitor of the reaction, even at a high ratio of xylitol to l-xylulose the inhibition did not reach 50%.  相似文献   

10.
Oxidation of l-serine and l-threonine by a silver(III) complex anion, [Ag(HIO(6))(2)](5-), has been studied in aqueous alkaline medium. The oxidation products of the amino acids have been identified as ammonia, glyoxylic acid and aldehyde (formaldehyde for serine and acetaldehyde for threonine). Kinetics of the oxidation reactions has been followed by the conventional spectrophotometry in the temperature range of 20.0-35.0 degrees C and the reactions display an overall second-order behavior: first-order with respect to both Ag(III) and the amino acids. Analysis of influences of [OH(-)] and [periodate] on the second-order rate constants k' reveals an empirical rate expression: k(')=(k(a)+k(b)[OH(-)])K(1)/([H(2)IO(6)(3-)](e)+K(1)), where [H(2)IO(6)(3-)](e) is equilibrium concentration of periodate, and where k(a)=6.1+/-0.5M(-1)s(-1), k(b)=264+/-6M(-2)s(-1), and K(1)=(6.5+/-1.3)x10(-4)M for serine and k(a)=12.6+/-1.7M(-1)s(-1), k(b)=(5.5+/-0.2)x10(2)M(-2)s(-1), and K(1)=(6.2+/-1.5)x10(-4)M for threonine at 25.0 degrees C and ionic strength of 0.30M. Activation parameters associated with k(a) and k(b) have also been derived. A reaction mechanism is proposed to involve two pre-equilibria, leading to formation of an Ag(III)-periodato-amino acid ternary complex. The ternary complex undergoes a two-electron transfer from the coordinated amino acid to the metal center via two parallel pathways: one pathway is spontaneous and the other is assisted by a hydroxide ion. Potential applications of the Ag(III) complex as a reagent for modifications of peptides and proteins are implicated.  相似文献   

11.
12.
Pseudomonas cichoriiid-tagatose 3-epimerase (P. cichoriid-TE) can efficiently catalyze the epimerization of not only d-tagatose to d-sorbose, but also d-fructose to d-psicose, and is used for the production of d-psicose from d-fructose. The crystal structures of P. cichoriid-TE alone and in complexes with d-tagatose and d-fructose were determined at resolutions of 1.79, 2.28, and 2.06 Å, respectively. A subunit of P. cichoriid-TE adopts a (β/α)8 barrel structure, and a metal ion (Mn2+) found in the active site is coordinated by Glu152, Asp185, His211, and Glu246 at the end of the β-barrel. P. cichoriid-TE forms a stable dimer to give a favorable accessible surface for substrate binding on the front side of the dimer. The simulated omit map indicates that O2 and O3 of d-tagatose and/or d-fructose coordinate Mn2+, and that C3-O3 is located between carboxyl groups of Glu152 and Glu246, supporting the previously proposed mechanism of deprotonation/protonation at C3 by two Glu residues. Although the electron density is poor at the 4-, 5-, and 6-positions of the substrates, substrate-enzyme interactions can be deduced from the significant electron density at O6. The O6 possibly interacts with Cys66 via hydrogen bonding, whereas O4 and O5 in d-tagatose and O4 in d-fructose do not undergo hydrogen bonding to the enzyme and are in a hydrophobic environment created by Phe7, Trp15, Trp113, and Phe248. Due to the lack of specific interactions between the enzyme and its substrates at the 4- and 5-positions, P. cichoriid-TE loosely recognizes substrates in this region, allowing it to efficiently catalyze the epimerization of d-tagatose and d-fructose (C4 epimer of d-tagatose) as well. Furthermore, a C3-O3 proton-exchange mechanism for P. cichoriid-TE is suggested by X-ray structural analysis, providing a clear explanation for the regulation of the ionization state of Glu152 and Glu246.  相似文献   

13.
The araA gene encoding L-arabinose isomerase (AI) from the thermoacidophilic bacterium Alicyclobacillus acidocaldarius was cloned, sequenced, and expressed in Escherichia coli. Analysis of the sequence revealed that the open reading frame of the araA gene consists of 1,491 bp that encodes a protein of 497 amino acid residues with a calculated molecular mass of 56,043 Da. Comparison of the deduced amino acid sequence of A. acidocaldarius AI (AAAI) with other AIs demonstrated that AAAI has 97% and 66% identities (99% and 83% similarities) to Geobacillus stearothermophilus AI (GSAI) and Bacillus halodurans AI (BHAI), respectively. The recombinant AAAI was purified to homogeneity by heat treatment, ion-exchange chromatography, and gel filtration. The purified enzyme showed maximal activity at pH 6.0 to 6.5 and 65 degrees C under the assay conditions used, and it required divalent cations such as Mn2+, Co2+, and Mg2+ for its activity. The isoelectric point (pI) of the enzyme was about 5.0 (calculated pI of 5.5). The apparent Km values of the recombinant AAAI for L-arabinose and D-galactose were 48.0 mM (Vmax, 35.5 U/mg) and 129 mM (Vmax, 7.5 U/mg), respectively, at pH 6 and 65 degrees C. Interestingly, although the biochemical properties of AAAI are quite similar to those of GSAI and BHAI, the three AIs from A. acidocaldarius (pH 6), G. stearothermophilus (pH 7), and B. halodurans (pH 8) exhibited different pH activity profiles. Based on alignment of the amino acid sequences of these homologous AIs, we propose that the Lys-269 residue of AAAI may be responsible for the ability of the enzyme to act at low pH. To verify the role of Lys-269, we prepared the mutants AAAI-K269E and BHAI-E268K by site-directed mutagenesis and compared their kinetic parameters with those of wild-type AIs at various pHs. The pH optima of both AAAI-K269E and BHAI-E268K were rendered by 1.0 units (pH 6 to 7 and 8 to 7, respectively) compared to the wild-type enzymes. In addition, the catalytic efficiency (kcat/Km) of each mutant at different pHs was significantly affected by an increase or decrease in Vmax. From these results, we propose that the position corresponding to the Lys-269 residue of AAAI could play an important role in the determination of the pH optima of homologous AIs.  相似文献   

14.
D-amino acid amidase (DAA) from Ochrobactrum anthropi SV3, which catalyzes the stereospecific hydrolysis of D-amino acid amides to yield the D-amino acid and ammonia, has attracted increasing attention as a catalyst for the stereospecific production of D-amino acids. In order to clarify the structure-function relationships of DAA, the crystal structures of native DAA, and of the D-phenylalanine/DAA complex, were determined at 2.1 and at 2.4 A resolution, respectively. Both crystals contain six subunits (A-F) in the asymmetric unit. The fold of DAA is similar to that of the penicillin-recognizing proteins, especially D-alanyl-D-alanine-carboxypeptidase from Streptomyces R61, and class C beta-lactamase from Enterobacter cloacae strain GC1. The catalytic residues of DAA and the nucleophilic water molecule for deacylation were assigned based on these structures. DAA has a flexible Omega-loop, similar to class C beta-lactamase. DAA forms a pseudo acyl-enzyme intermediate between Ser60 O(gamma) and the carbonyl moiety of d-phenylalanine in subunits A, B, C, D, and E, but not in subunit F. The difference between subunit F and the other subunits (A, B, C, D and E) might be attributed to the order/disorder structure of the Omega-loop: the structure of this loop cannot assigned in subunit F. Deacylation of subunit F may be facilitated by the relative movement of deprotonated His307 toward Tyr149. His307 N(epsilon2) extracts the proton from Tyr149 O(eta), then Tyr149 O(eta) attacks a nucleophilic water molecule as a general base. Gln214 on the Omega-loop is essential for forming a network of water molecules that contains the nucleophilic water needed for deacylation. Although peptidase activity is found in almost all penicillin-recognizing proteins, DAA lacks peptidase activity. The lack of transpeptidase and carboxypeptidase activities may be attributed to steric hindrance of the substrate-binding pocket by a loop comprised of residues 278-290 and the Omega-loop.  相似文献   

15.
The high molecular weight of chitosan, which results in a poor solubility at neutral pH values and high viscosity aqueous solutions, limits its potential uses in the fields of food, health and agriculture. However, most of these limitations are overcome by chitosan oligosaccharides obtained by enzymatic hydrolysis of the polymer. Several commercial enzymes with different original specificities were assayed for their ability to hydrolyze a 93% deacetylation degree chitosan and compared with a chitosanase. According to the patterns of viscosity decrease and reducing end formation, three enzymes--cellulase, pepsin and lipase A--were found to be particularly suitable for hydrolyzing chitosan at a level comparable to that achieved by chitosanase. Unlike the appreciable levels of both 2-amino-2-deoxy-D-glucose and 2-acetamido-2-deoxy-D-glucose monomers released from chitosan by the other enzymes after a 20h-hydrolysis (4.6-9.1% of the total product weight), no monomer could be detected following pepsin cleavage. As a result, pepsin produced a higher yield of chitosan oligosaccharides than the other enzymes: 52% versus as much as 46%, respectively. Low molecular weight chitosans accounted for the remaining 48% of hydrolysis products. The calculated average polymerization degree of the products released by pepsin was around 16 units after 20h of hydrolysis. This product pattern and yield are proposed to be related to the bond cleavage specificity of pepsin and the high deacetylation degree of chitosan used as substrate. The optimal reaction conditions for hydrolysis of chitosan by pepsin were 40 degrees C and pH 4.5, and an enzyme/substrate ratio of 1:100 (w/w) for reactions longer than 1h.  相似文献   

16.
17.
Biosynthesis of lysine and meso-diaminopimelic acid in bacteria provides essential components for protein synthesis and construction of the bacterial peptidoglycan cell wall. The dapE operon enzymes synthesize both meso-diaminopimelic acid and lysine and, therefore, represent potential targets for novel antibacterials. The dapE-encoded N-succinyl-l,l-diaminopimelic acid desuccinylase functions in a late step of the pathway and converts N-succinyl-l,l-diaminopimelic acid to l,l-diaminopimelic acid and succinate. Deletion of the dapE gene is lethal to Helicobacter pylori and Mycobacterium smegmatis, indicating that DapE's are essential for cell growth and proliferation. Since there are no similar pathways in humans, inhibitors that target DapE may have selective toxicity against only bacteria. A major limitation in developing antimicrobial agents that target DapE has been the lack of structural information. Herein, we report the high-resolution X-ray crystal structures of the DapE from Haemophilus influenzae with one and two zinc ions bound in the active site, respectively. These two forms show different activity. Based on these newly determined structures, we propose a revised catalytic mechanism of peptide bond cleavage by DapE enzymes. These structures provide important insight into catalytic mechanism of DapE enzymes as well as a structural foundation that is critical for the rational design of DapE inhibitors.  相似文献   

18.
Fe-only hydrogenases contain a di-iron active site complex, in which the two Fe atoms have carbon monoxide and cyanide ligands and are linked together by a putative di(thiomethyl)amine molecule. We have cloned, purified and characterized the HydE and HydG proteins, thought to be involved in the biosynthesis of this peculiar metal site, from the thermophilic organism Thermotoga maritima. The HydE protein anaerobically reconstituted with iron and sulfide binds two [4Fe-4S] clusters, as characterized by UV and EPR spectroscopy. The HydG protein binds one [4Fe-4S] cluster, and probably an additional one. Both enzymes are able to reductively cleave S-adenosylmethionine (SAM) when reduced by dithionite, confirming that they are Radical-SAM enzymes.  相似文献   

19.
A putative long-chain mannitol-1-phosphate 5-dehydrogenase from Aspergillus fumigatus (AfM1PDH) was overexpressed in Escherichia coli to a level of about 50% of total intracellular protein. The purified recombinant protein was a approximately 40-kDa monomer in solution and displayed the predicted enzymatic function, catalyzing NAD(H)-dependent interconversion of d-mannitol 1-phosphate and d-fructose 6-phosphate with a specific reductase activity of 170 U/mg at pH 7.1 and 25 degrees C. NADP(H) showed a marginal activity. Hydrogen transfer from formate to d-fructose 6-phosphate, mediated by NAD(H) and catalyzed by a coupled enzyme system of purified Candida boidinii formate dehydrogenase and AfM1PDH, was used for the preparative synthesis of d-mannitol 1-phosphate or, by applying an analogous procedure using deuterio formate, the 5-[2H] derivative thereof. Following the precipitation of d-mannitol 1-phosphate as barium salt, pure product (>95% by HPLC and NMR) was obtained in isolated yields of about 90%, based on 200 mM of d-fructose 6-phosphate employed in the reaction. In situ proton NMR studies of enzymatic oxidation of d-5-[2H]-mannitol 1-phosphate demonstrated that AfM1PDH was stereospecific for transferring the deuterium to NAD+, producing (4S)-[2H]-NADH. Comparison of maximum initial rates for NAD+-dependent oxidation of protio and deuterio forms of D-mannitol 1-phosphate at pH 7.1 and 25 degrees C revealed a primary kinetic isotope effect of 2.9+/-0.2, suggesting that the hydride transfer was strongly rate-determining for the overall enzymatic reaction under these conditions.  相似文献   

20.
Following a report of variations in the lipopolysaccharide (LPS) structure of Yersinia pestis at mammalian (37 degrees C) and flea (25 degrees C) temperatures, a number of changes to the LPS structure were observed when the bacterium was cultivated at a temperature of winter-hibernating rodents (6 degrees C). In addition to one of the known Y. pestis LPS types, LPS of a new type was isolated from Y. pestis KM218 grown at 6 degrees C. The core of the latter differs in: (i) replacement of terminal galactose with terminal d-glycero-d-manno-heptose; (ii) phosphorylation of terminal oct-2-ulosonic acid with phosphoethanolamine; (iii) a lower content of GlcNAc, and; (iv) the absence of glycine; lipid A differs in the lack of any 4-amino-4-deoxyarabinose and presumably partial (di)oxygenation of a fatty acid(s). The data obtained suggest that cold temperature switches on an alternative mechanism of control of the synthesis of Y. pestis LPS.  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