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.
The acetate kinase from the Antarctic psychrophilic Shewanella sp. AS-11 (SAK) has a significantly higher catalytic efficiency at low temperatures when compared with that from mesophilic Escherichia coli K-12 (EAK). To examine the stability and conformational flexibility of SAK and EAK, steady state intrinsic fluorescence studies were performed. EAK contains only one Trp at a position 46, while SAK contains two Trps at positions 46 and 388. From the fluorescence emission spectra, quenching with acrylamide, Cs+ and I at different temperatures and denaturation with guanidine-HCl, it was revealed that the SAK bears more flexible and unstable structure than that of EAK. Substrate-induced conformational changes reflect that SAK reached transition state through more conformational changes than EAK. In combination of our thermodynamic studies on the enzymatic reaction and present research findings, it can be concluded that these structural features of SAK may contribute to its high catalytic efficiency at low temperatures.  相似文献   

2.
We investigated the relationship between Escherichia coli flagellar expression and the regulation of acetyl phosphate synthesis and degradation. Using cells either wild type for acetyl phosphate metabolism or defective for phosphotransacetylase or acetate kinase, or both, we measured flagellar expression and the intracellular concentration of acetyl phosphate relative to growth phase and temperature. Under the conditions tested, we found that elevated levels of acetyl phosphate corresponded to inhibition of flagellar synthesis. To extend these observations, we measured the intracellular concentration of acetyl-CoA, the level of expression from the pta and ackA promoters, and the activities of phosphotransacetylase and acetate kinase derived from cell lysates. Relative to increasing culture density, acetyl-CoA levels and expression from both the pta and ackA promoters decreased. Relative to Increasing temperature, expression from the ackA promoter decreased and phosphotransacetylase activity increased. In contrast, temperature had little or no effect on either acetate kinase activity or expression from the pta promoter. We propose that cells regulate intracellular acetyl phosphate concentrations relative to growth phase and temperature by modulating the availability of acetyl-CoA, the expression of ackA, and the activity of phosphotransacetylase.  相似文献   

3.
Acetate kinase, a member of the acetate and sugar kinase-Hsp70-actin (ASKHA) enzyme superfamily1-5, is responsible for the reversible phosphorylation of acetate to acetyl phosphate utilizing ATP as a substrate. Acetate kinases are ubiquitous in the Bacteria, found in one genus of Archaea, and are also present in microbes of the Eukarya6. The most well characterized acetate kinase is that from the methane-producing archaeon Methanosarcina thermophila7-14. An acetate kinase which can only utilize PPi but not ATP in the acetyl phosphate-forming direction has been isolated from Entamoeba histolytica, the causative agent of amoebic dysentery, and has thus far only been found in this genus15,16.In the direction of acetyl phosphate formation, acetate kinase activity is typically measured using the hydroxamate assay, first described by Lipmann17-20, a coupled assay in which conversion of ATP to ADP is coupled to oxidation of NADH to NAD+ by the enzymes pyruvate kinase and lactate dehydrogenase21,22, or an assay measuring release of inorganic phosphate after reaction of the acetyl phosphate product with hydroxylamine23. Activity in the opposite, acetate-forming direction is measured by coupling ATP formation from ADP to the reduction of NADP+ to NADPH by the enzymes hexokinase and glucose 6-phosphate dehydrogenase24.Here we describe a method for the detection of acetate kinase activity in the direction of acetate formation that does not require coupling enzymes, but is instead based on direct determination of acetyl phosphate consumption. After the enzymatic reaction, remaining acetyl phosphate is converted to a ferric hydroxamate complex that can be measured spectrophotometrically, as for the hydroxamate assay. Thus, unlike the standard coupled assay for this direction that is dependent on the production of ATP from ADP, this direct assay can be used for acetate kinases that produce ATP or PPi.  相似文献   

4.
Acetate kinase (ACK) converts acetyl phosphate to acetate along with the generation of ATP in the pathway for mixed-acid fermentation in Lactococcus lactis. The reverse reaction yields acetyl phosphate for assimilation purposes. Remarkably, L. lactis has two ACK isozymes, and the corresponding genes are present in an operon. We purified both enzymes (AckA1 and AckA2) from L. lactis MG1363 and determined their oligomeric state, specific activities, and allosteric regulation. Both proteins form homodimeric complexes, as shown by size exclusion chromatography and static light-scattering measurements. The turnover number of AckA1 is about an order of magnitude higher than that of AckA2 for the reaction in either direction. The Km values for acetyl phosphate, ATP, and ADP are similar for both enzymes. However, AckA2 has a higher affinity for acetate than does AckA1, suggesting an important role under acetate-limiting conditions despite the lower activity. Fructose-1,6-bisphosphate, glyceraldehyde-3-phosphate, and phospho-enol-pyruvate inhibit the activities of AckA1 and AckA2 to different extents. The allosteric regulation of AckA1 and AckA2 and the pool sizes of the glycolytic intermediates are consistent with a switch from homolactic to mixed-acid fermentation upon slowing of the growth rate.  相似文献   

5.
Acetate kinase (ACK) catalyzes the reversible synthesis of acetyl phosphate by transfer of the γ-phosphate of ATP to acetate. Here we report the first biochemical and kinetic characterization of a eukaryotic ACK, that from the protist Entamoeba histolytica. Our characterization revealed that this protist ACK is the only known member of the ASKHA structural superfamily, which includes acetate kinase, hexokinase, and other sugar kinases, to utilize inorganic pyrophosphate (PPi)/inorganic phosphate (Pi) as the sole phosphoryl donor/acceptor. Detection of ACK activity in E. histolytica cell extracts in the direction of acetate/PPi formation but not in the direction of acetyl phosphate/Pi formation suggests that the physiological direction of the reaction is toward acetate/PPi production. Kinetic parameters determined for each direction of the reaction are consistent with this observation. The E. histolytica PPi-forming ACK follows a sequential mechanism, supporting a direct in-line phosphoryl transfer mechanism as previously reported for the well-characterized Methanosarcina thermophila ATP-dependent ACK. Characterizations of enzyme variants altered in the putative acetate/acetyl phosphate binding pocket suggested that acetyl phosphate binding is not mediated solely through a hydrophobic interaction but also through the phosphoryl group, as for the M. thermophila ACK. However, there are key differences in the roles of certain active site residues between the two enzymes. The absence of known ACK partner enzymes raises the possibility that ACK is part of a novel pathway in Entamoeba.  相似文献   

6.
Acetate kinase, a member of the ASKHA (Acetate and Sugar Kinases, Hsp70, Actin) phosphotransferase superfamily is a central enzyme in prokaryotic carbon and energy metabolism. Recently extensive structural and biochemical studies of acetate kinase and related carboxylate kinases have been conducted. Analysis of the kinetic properties of wild-type and mutant enzymes has been impeded by the nature of the current assays for acetate kinase activity. These assays have the disadvantages of being either discontinuous or insensitive or of utilizing compounds that interfere with activity measurements. We have developed a novel continuous assay that depends on the purine nucleoside phosphorylase-based spectroscopic measurement of the inorganic phosphate generated by hydroxylaminolysis of one of the products of the acetate kinase reaction, acetyl phosphate. This assay has enabled a determination of the kinetic parameters of the Thermotoga maritima acetate kinase that indicates a lower Km for acetate than previously published.  相似文献   

7.
An enzyme from Entamoeba histolytica catalyzes the formation of acetyl phosphate and orthophosphate from acetate and inorganic pyrophosphate (PPi), but it displays much greater activity in the direction of acetate formation. It has been purified 40-fold and separated from interfering enzyme activities by chromatography. Its reaction products have been quantitatively established. ATP cannot replace PPi as phosphoryl donor in the direction of acetyl phosphate formation nor will any common nucleoside diphosphate replace orthophosphate as phosphoryl acceptor in the direction of acetate formation. The trivial name proposed for the new enzyme is acetate kinase (PPi).  相似文献   

8.
Acetate kinase catalyzes the reversible phosphorylation of acetate (CH3COO + ATPCH3CO2PO32− + ADP). A mechanism which involves a covalent phosphoryl-enzyme intermediate has been proposed, and chemical modification studies of the enzyme from Escherichia coli indicate an unspecified glutamate residue is phosphorylated (J. A. Todhunter and D. L. Purich, Biochem. Biophys. Res. Commun. 60:273–280, 1974). Alignment of the amino acid sequences for the acetate kinases from E. coli (Bacteria domain), Methanosarcina thermophila (Archaea domain), and four other phylogenetically divergent microbes revealed high identity which included five glutamates. These glutamates were replaced in the M. thermophila enzyme to determine if any are essential for catalysis. The histidine-tagged altered enzymes were produced in E. coli and purified to electrophoretic homogeneity by metal affinity chromatography. Replacements of E384 resulted in either undetectable or extremely low kinase activity, suggesting E384 is essential for catalysis which supports the proposed mechanism. Replacement of E385 influenced the Km values for acetate and ATP with only moderate decreases in kcat, which suggests that this residue is involved in substrate binding but not catalysis. The unaltered acetate kinase was not inactivated by N-ethylmaleimide; however, replacement of E385 with cysteine conferred sensitivity to N-ethylmaleimide which was prevented by preincubation with acetate, acetyl phosphate, ATP, or ADP, suggesting that E385 is located near the active site. Replacement of E97 decreased the Km value for acetate but not ATP, suggesting this residue is involved in binding acetate. Replacement of either E32 or E334 had no significant effects on the kinetic constants, which indicates that neither residue is essential for catalysis or significantly influences the binding of acetate or ATP.  相似文献   

9.
Non-growing cells of Escherichia coli O157:H7 and K-12 that were incubated anaerobically in sodium phosphate buffer at pH 6.5 consumed glucose at a rate of approximately 8 μmol·(mg protein)−1·h−1 and had intracellular pH values of 7.3 and 7.5, respectively. The uncoupler, carbonylcyanide-m-chlorophenylhydrazone (CCCP), caused a marked decrease in intracellular pH, ATP and potassium of both strains. Low concentrations of CCCP stimulated glucose consumption rate, but higher concentrations were inhibitory. Acetate also caused a decrease in intracellular pH, but it never caused a large decrease in glucose consumption rate. Acetate decreased the intracellular ATP of E. coli K-12, but it had no effect on the ATP of O157:H7. Acetate had no effect on the intracellular potassium of E. coli O157:H7, and acetate-treated K-12 cells had even more potassium than untreated controls. Based on these results, acetate and CCCP appear to have different effects on E. coli. The comparison of E. coli O157:H7 and K-12 indicated that intracellular pH, acetate accumulation and intracellular potassium were related. E. coli K-12 maintained a higher intracellular pH than O157:H7, accumulated more acetate and had a greater intracellular potassium.  相似文献   

10.
The anaerobic hyperthermophilic archaea Desulfurococcus amylolyticus, Hyperthermus butylicus, Thermococcus celer, Pyrococcus woesei, the hyperthermophilic bacteria Thermotoga maritima and Clostridium thermohydrosulfuricum and the aerobic mesophilic archaeon Halobacterium saccharovorum were grown either on complex media, on sugars or on pyruvate as carbon and energy sources. During growth acetate was formed as fermentation product by all organisms. The enzymes involved in acetyl-CoA formation from pyruvate and in acetate formation from acetyl-CoA were investigated:
  1. Cell extracts of all species, both archaea and bacteria, catalyzed the coenzyme A-dependent oxidative decarboxylation of pyruvate with viologen dyes or with Clostridium pasteurianum ferredoxin as electron acceptors indicating a pyruvate: ferredoxin oxidoreductase to be operative in acetyl-CoA formation from pyruvate.
  2. Cell extracts of all archaeal species, both hyperthermophiles (D. amylolyticus, H. butylicus, T. celer, P. woesei) and the mesophile H. saccharovorum, contained an acetyl-CoA synthetase (ADP forming), which catalyzes both acetate formation from acetyl-CoA and ATP synthesis from ADP and phosphate (Pi): Acetyl-CoA+ADP+Pi?Acetate + ATP+CoA. Phosphate acetyltransferase and acetate kinase could not be detected.
  3. Cell extracts of the hyperthermophilic (eu)bacteria T. maritima and C. thermohydrosulfuricum contained phosphate acetyltransferase and acetate kinase rather than acetyl-CoA synthetase (ADP forming).
These data indicate that acetyl-CoA synthetase (ADP forming) represents a typical archaeal property rather than an enzyme specific for hyperthermophiles. It is proposed that in all acetate forming archaea the formation of acetate and of ATP from acetyl-CoA, ADP and Pi are catalyzed by acetyl-CoA synthetase (ADP forming), whereas in all acetate forming (eu)bacteria these reactions are catalyzed by two enzymes, phosphate acetyltransferase and acetate kinase.  相似文献   

11.
Procedures are described for isolating highly purified porcine liver pyruvate and α-ketoglutarate dehydrogenase complexes. Rabbit serum stabilized these enzyme complexes in mitochondrial extracts, apparently by inhibiting lysosomal proteases. The complexes were purified by a three-step procedure involving fractionation with polyethylene glycol, pelleting through 12.5% sucrose, and a second fractionation under altered conditions with polyethylene glycol. Sedimentation equilibrium studies gave a molecular weight of 7.2 × 106 for the liver pyruvate dehydrogenase complex. Kinetic parameters are presented for the reaction catalyzed by the pyruvate dehydrogenase complex and for the regulatory reactions catalyzed by the pyruvate dehydrogenase kinase and pyruvate dehydrogenase phosphatase. For the overall catalytic reaction, the competitive Ki to Km ratio for NADH versus NAD+ and acetyl CoA versus CoA were 4.7 and 5.2, respectively. Near maximal stimulations of pyruvate dehydrogenase kinase by NADH and acetyl CoA were observed at NADH:NAD+ and acetyl CoA:CoA ratios of 0.15 and 0.5, respectively. The much lower ratios required for enhanced inactivation of the complex by pyruvate dehydrogenase kinase than for product inhibition indicate that the level of activity of the regulatory enzyme is not directly determined by the relative affinity of substrates and products of catalytic sites in the pyruvate dehydrogenase complex. In the pyruvate dehydrogenase kinase reaction, K+ and NH+4 decreased the Km for ATP and the competitive inhibition constants for ADP and (β,γ-methylene)adenosine triphosphate. Thiamine pyrophosphate strongly inhibited kinase activity. A high concentration of ADP did not alter the degree of inhibition by thiamine pyrophosphate nor did it increase the concentration of thiamine pyrophosphate required for half-maximal inhibition.  相似文献   

12.
Acetate-mediated growth inhibition of Escherichia coli has been found to be a consequence of the accumulation of homocysteine, the substrate of the cobalamin-independent methionine synthase (MetE) that catalyzes the final step of methionine biosynthesis. To improve the acetate resistance of E. coli, we randomly mutated the MetE enzyme and isolated a mutant enzyme, designated MetE-214 (V39A, R46C, T106I, and K713E), that conferred accelerated growth in the E. coli K-12 WE strain in the presence of acetate. Additionally, replacement of cysteine 645, which is a unique site of oxidation in the MetE protein, with alanine improved acetate tolerance, and introduction of the C645A mutation into the MetE-214 mutant enzyme resulted in the highest growth rate in acetate-treated E. coli cells among three mutant MetE proteins. E. coli WE strains harboring acetate-tolerant MetE mutants were less inhibited by homocysteine in l-isoleucine-enriched medium. Furthermore, the acetate-tolerant MetE mutants stimulated the growth of the host strain at elevated temperatures (44 and 45°C). Unexpectedly, the mutant MetE enzymes displayed a reduced melting temperature (Tm) but an enhanced in vivo stability. Thus, we demonstrate improved E. coli growth in the presence of acetate or at elevated temperatures solely due to mutations in the MetE enzyme. Furthermore, when an E. coli WE strain carrying the MetE mutant was combined with a previously found MetA (homoserine o-succinyltransferase) mutant enzyme, the MetA/MetE strain was found to grow at 45°C, a nonpermissive growth temperature for E. coli in defined medium, with a similar growth rate as if it were supplemented by l-methionine.  相似文献   

13.
14.
The surface properties at the air/water interface and the interaction of branched chain polymeric polypeptides with a general formula poly[Lys-(DL -Alam-X1)], where X = Π (AK), Ser (SAK), or Glu (EAK), with phospholipids were investigated. Polylysine derivatives with polycationic (SAK, AK) or amphoteric (EAK) were capable to spread and form stable monomolecular layers. The stability of monolayers at the air/water interface was dependent on the side-chain terminal amino acid residue of polymers and can be described by SAK < AK < EAK order. The area per amino acid residue values calculated from compression isotherms were in the same range as compared to those of linear poly-α-amino acids and proteins. Moreover, these polymers interact with phospholipid monomolecular layers composed of dipalmitoyl phosphatidyl choline (DPPC) or DPPC/PG (PG: phosphatidyl glycerol; 95/5, mol/mol). Data obtained from compression isotherms of phospholipids spread on aqueous polymer solutions at different initial surface pressure indicated that insertion into lipid monolayers for SAK or AK is more pronounced than for EAK. The interaction between branched polypeptides and phospholipid membranes was further investigated using lipid bilayers with DPPC/PG and fluorescent probes located either at the polar surface [1-(4-trimethylammonium-phenyl)-6-phenyl-1,3,5-hexatriene (TMA-DPH) sodium anilino naphthalene sulfonate (ANS)] or within the hydrophobic core (DPH) of the liposome. Changes in fluorescence intensity and in polarization were observed when TMA-DPH or ANS, but not DPH were used. Comparative data also indicate that all three polymers interact only with the outer surface of the bilayer, but even the most marked penetration of polycationic polypeptide (SAK) did not result in alteration of the ordered state of the alkyl chains in the bilayer. Taken together, data obtained from mono- or bilayer experiments suggest that the interaction between branched polymers and phospholipids are highly dependent on the charge properties (Ser vs Glu) and on the identity (Ser vs Ala) of side-chain terminating amino acids. The binding of polymers to the model membranes could be mainly driven by electrostatic forces, but the significant role of hydrophilic properties in case of SAK cannot be excluded. © 1998 John Wiley & Sons, Inc. Biopoly 46: 169–179, 1998  相似文献   

15.
A process of glucose-6-phosphate (G-6-P) production coupled with an adenosine triphosphate (ATP) regeneration system was constructed that utilized acetyl phosphate (ACP) via acetate kinase (ACKase). The genes glk and ack from Escherichia coli K12 were amplified and cloned into pET-28a(+), then transformed into E. coli BL21 (DE3) and the recombinant strains were named pGLK and pACK respectively. Glucokinase (glkase) in pGLK and ACKase in pACK were both overexpressed in soluble form. G-6-P was efficiently produced from glucose and ACP using a very small amount of ATP. The conversion yield was greater than 97 % when the reaction solution containing 10 mM glucose, 20 mM ACP-Na2, 0.5 mM ATP, 5 mM Mg2+, 50 mM potassium phosphate buffer (pH 7.0), 4.856 U glkase and 3.632 U ACKase were put into 37 °C water bath for 1 h.  相似文献   

16.
Phosphotransacetylase (EC 2.3.1.8) was detected in cell-free crude extracts of starch-fermenting eucaryotic green algae. The enzyme was purified from autotrophically grown Chlorogonium elongatum. The purified enzyme fraction, after affinity chromatography, shows a single protein band upon acrylamide gel electrophoresis and has a molecular weight of 280 000. It consists of six subunits of identical molecular weight (44 000). The pH and temperature optima for the eucaryotic phosphotransacetylase are 7.6 and 28°C, respectively. The Km values at 25°C (pH 7.6) for acetyl-CoA and phosphate are 0.078 mM and 5.440 mM, respectively, and in the reverse reaction (acetyl-CoA synthesis) for CoA and acetyl phosphate 0.093 mM and 0.310 mM, respectively. The maximum velocity of the forward reaction was 1627 nkat/mg protein and of the reverse reaction 8582 nkat/mg protein. The activity of the eucaryotic phosphotransacetylase strictly depends on the presence of univalent cations (ammonium, Ka = 9 mM; potassium, Ka = 12.5 mM). Inactivation studies with iodoacetamide and iodoacetic acid revealed the presence of an essential sulphhydryl group at the catalytic site. Arsenolytic and product inhibition studies indicate a rapid equilibrium random bi-bi reaction mechanism for the enzyme from C. elongatum. The control of the enzyme activity in the forward reaction by both pyruvate and NADH gives evidence for a physiological function of phosphotransacetylase in anaerobic energy metabolism of eucaryotic green algae rather than in aerobic acetate activation.  相似文献   

17.
Five recombinant Escherichia coli extracts harboring overexpressed galactokinase, galactose-1-phosphate uridyltransferase, UDP-glucose pyrophophorylase, UMP kinase, and acetate kinase (AK) were utilized for the production of UDP-galactose (UDP-Gal). We analyzed the parameters which limit the yield of UDP-Gal in the reaction, and the reaction was optimized by increasing the concentration of AK. AK was used for the ATP regeneration as well as the conversion of UDP to UTP. The activities of four overexpressed enzymes were identically fixed, and then we increased the activity of AK to 20 times higher than others. The extracts catalyzed the production of UDP-Gal from UMP (10 mM), galactose (12 mM), ATP (1 mM), and acetyl phosphate (40 mM). As the result of the reaction, the conversion yield of UDP-Gal reached to 95% from 10 mM UMP.  相似文献   

18.
Escherichia coli strains overproducing the response regulator CheY respond to acetate by increasing their clockwise bias of flagellar rotation, even when they lack other chemotaxis proteins. With acetate metabolism mutants, we demonstrate that both acetate kinase and acetyl coenzyme A synthetase are involved in this response. Thus, a response was observed when one of these enzymes was missing but not when both were absent.  相似文献   

19.
Acetate kinase (EC 2.7.2.1) was purified from Acholeplasma laidlawii cytoplasm by a combination of ammonium sulfate fractionation, gel filtration, diethylaminoethyl-cellulose chromatography, and affinity chromatography on 8-(6-aminohexylamino)-adenosine 5'-triphosphate conjugated to Sepharose 4B. The enzyme was composed of polypeptide chains of about 50,000 molecular weight as estimated from sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Under nondenaturating conditions, apparent molecular weights between 64,000 and 130,000 were obtained, depending upon mainly the ionic strength of the test solution. The enzyme had a narrow specificity for phosphate acceptor acids, whereas both purine and pyrimidine nucleoside triphosphates were suitable phosphate donors. Na(+) and K(+) inhibited both acetyl phosphate and adenosine 5'-triphosphate synthesis, and the latter was also inhibited by high concentrations of adenosine 5'-diphosphate and acetyl phosphate. This substrate inhibition was partially abolished by 0.5 M NaCl. The enzyme catalyzed the independent adenosine 5'-diphosphate<-->adenosine 5'-triphosphate and acetate<-->acetyl phosphate exchanges. The rate of the latter was enhanced by the addition of cosubstrate Mg(2+)-adenosine 5'-triphosphate. The high affinity for substrates, except for acetate, indicated that under physiological conditions the direction of the enzymic reaction favors adenosine 5'-triphosphate synthesis. Thus, a mechanism for adenosine 5'-triphosphate generation in mycoplasmas is suggested.  相似文献   

20.
Carbon sources that can be converted to acetate were added to the growth medium of Escherichia coli wild-type cells. Cells responded with an increased cell division rate. The addition of acetate also caused a decreased synthesis of flagella. Mutants in phosphotransacetylase, which are incapable of synthesizing acetyl phosphate, and mutants in the osmoregulator OmpR divided at a lower rate than did wild-type cells. The mutants did not increase their cell division rate upon the addition of serine, as observed for wild-type cells. These data are consistent with the idea that the previously described effect of serine upon the cell division rate is mediated by acetyl phosphate and phosphorylation of OmpR. Received: 23 March 1998 / Accepted: 8 May 1998  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