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 390 毫秒
1.
The genetic operon for propionic acid degradation in Salmonella enterica serovar Typhimurium contains an open reading frame designated prpE which encodes a propionyl coenzyme A (propionyl-CoA) synthetase (A. R. Horswill and J. C. Escalante-Semerena, Microbiology 145:1381-1388, 1999). In this paper we report the cloning of prpE by PCR, its overexpression in Escherichia coli, and the substrate specificity of the enzyme. When propionate was utilized as the substrate for PrpE, a K(m) of 50 microM and a specific activity of 120 micromol. min(-1). mg(-1) were found at the saturating substrate concentration. PrpE also activated acetate, 3-hydroxypropionate (3HP), and butyrate to their corresponding coenzyme A esters but did so much less efficiently than propionate. When prpE was coexpressed with the polyhydroxyalkanoate (PHA) biosynthetic genes from Ralstonia eutropha in recombinant E. coli, a PHA copolymer containing 3HP units accumulated when 3HP was supplied with the growth medium. To compare the utility of acyl-CoA synthetases to that of an acyl-CoA transferase for PHA production, PHA-producing recombinant strains were constructed to coexpress the PHA biosynthetic genes with prpE, with acoE (an acetyl-CoA synthetase gene from R. eutropha [H. Priefert and A. Steinbüchel, J. Bacteriol. 174:6590-6599, 1992]), or with orfZ (an acetyl-CoA:4-hydroxybutyrate-CoA transferase gene from Clostridium propionicum [H. E. Valentin, S. Reiser, and K. J. Gruys, Biotechnol. Bioeng. 67:291-299, 2000]). Of the three enzymes, PrpE and OrfZ enabled similar levels of 3HP incorporation into PHA, whereas AcoE was significantly less effective in this capacity.  相似文献   

2.
In both animal and plant acyl elongation systems, it has been proposed that fatty acids are first activated to acyl-coenzyme A (CoA) before their elongation, and that the ATP dependence of fatty acid elongation is evidence of acyl-CoA synthetase involvement. However, because CoA is not supplied in standard fatty acid elongation assays, it is not clear if CoA-dependent acyl-CoA synthetase activity can provide levels of acyl-CoAs necessary to support typical rates of fatty acid elongation. Therefore, we examined the role of acyl-CoA synthetase in providing the primer for acyl elongation in leek (Allium porrum L.) epidermal microsomes and Brassica napus L. cv Reston oil bodies. As presented here, fatty acid elongation was independent of CoA and proceeded at maximum rates with CoA-free preparations of malonyl-CoA. We also showed that stearic acid ([1-14C]18:0)-CoA was synthesized from [1-14C]18:0 in the presence of CoA-free malonyl-CoA or acetyl-CoA, and that [1-14C]18:0-CoA synthesis under these conditions was ATP dependent. Furthermore, the appearance of [1-14C]18:0 in the acyl-CoA fraction was simultaneous with its appearance in phosphatidylcholine. These data, together with the s of a previous study (A. Hlousek-Radojcic, H. Imai, J.G. Jaworski [1995] Plant J 8: 803–809) showing that exogenous [14C]acyl-CoAs are diluted by a relatively large endogenous pool before they are elongated, strongly indicated that acyl-CoA synthetase did not play a direct role in fatty acid elongation, and that phosphatidylcholine or another glycerolipid was a more likely source of elongation primers than acyl-CoAs.  相似文献   

3.
The Escherichia coli XL1-blue strain was metabolically engineered to synthesize poly(3-hydroxybutyrate-co-3-hydroxyvalerate) [P(3HB-co-3HV)] through 2-ketobutyrate, which is generated via citramalate pathway, as a precursor for propionyl-CoA. Two different metabolic pathways were examined for the synthesis of propionyl-CoA from 2-ketobutyrate. The first pathway is composed of the Dickeya dadantii 3937 2-ketobutyrate oxidase or the E. coli pyruvate oxidase mutant (PoxB L253F V380A) for the conversion of 2-ketobutyrate into propionate and the Ralstonia eutropha propionyl-CoA synthetase (PrpE) or the E. coli acetyl-CoA:acetoacetyl-CoA transferase for further conversion of propionate into propionyl-CoA. The second pathway employs pyruvate formate lyase encoded by the E. coli tdcE gene or the Clostridium difficile pflB gene for the direct conversion of 2-ketobutyrate into propionyl-CoA. As the direct conversion of 2-ketobutyrate into propionyl-CoA could not support the efficient production of P(3HB-co-3HV) from glucose, the first metabolic pathway was further examined. When the recombinant E. coli XL1-blue strain equipped with citramalate pathway expressing the E. coli poxB L253F V380A gene and R. eutropha prpE gene together with the R. eutropha PHA biosynthesis genes was cultured in a chemically defined medium containing 20 g/L of glucose as a sole carbon source, P(3HB-co-2.3 mol% 3HV) was produced up to the polymer content of 61.7 wt.%. Moreover, the 3HV monomer fraction in P(3HB-co-3HV) could be increased up to 5.5 mol% by additional deletion of the prpC and scpC genes, which are responsible for the metabolism of propionyl-CoA in host strains.  相似文献   

4.
Polyhydroxyalkanoates (PHAs) are a class of carbon and energy storage polymers produced by numerous bacteria in response to environmental limitation. The type of polymer produced depends on the carbon sources available, the flexibility of the organism’s intermediary metabolism, and the substrate specificity of the PHA biosynthetic enzymes. Ralstonia eutropha produces both the homopolymer poly-β-hydroxybutyrate (PHB) and, when provided with the appropriate substrate, the copolymer poly(β-hydroxybutyrate-co-β-hydroxyvalerate) (PHBV). A required step in production of the hydroxyvalerate moiety of PHBV is the condensation of acetyl coenzyme A (acetyl-CoA) and propionyl-CoA to form β-ketovaleryl-CoA. This activity has generally been attributed to the β-ketothiolase encoded by R. eutropha phbA. However, we have determined that PhbA does not significantly contribute to catalyzing this condensation reaction. Here we report the cloning and genetic analysis of bktB, which encodes a β-ketothiolase from R. eutropha that is capable of forming β-ketovaleryl-CoA. Genetic analyses determined that BktB is the primary condensation enzyme leading to production of β-hydroxyvalerate derived from propionyl-CoA. We also report an additional β-ketothiolase, designated BktC, that probably serves as a secondary route toward β-hydroxyvalerate production.Polyhydroxyalkanoates (PHAs) are a class of naturally occurring polymers which serve as a carbon and energy reserve in numerous bacterial species. Ralstonia eutropha (formerly designated Alcaligenes eutrophus [41]) produces the homopolymer poly(β-hydroxybutyrate) (PHB) and, when provided with propionate in the feedstock, the copolymer poly(β-hydroxybutyrate-co-β-hydroxyvalerate) (PHBV). R. eutropha is used commercially to produce PHBV, which is a biodegradable thermoplastic.The PHB biosynthetic pathway requires three enzymatic activities: a β-ketothiolase (PhbA), an NADPH-dependent acetoacetyl coenzyme A (acetoacetyl-CoA) reductase (PhbB) and a PHB synthase (PhbC). The first step in production of the homopolymer PHB is catalyzed by β-ketothiolase which condenses two acetyl-CoA molecules to form acetoacetyl-CoA. Formation of the copolymer PHBV requires the additional condensation of acetyl-CoA with propionyl-CoA to form β-ketovaleryl-CoA (Fig. (Fig.1).1). Subsequently, the acetoacetyl-CoA and β-ketovaleryl-CoA are converted into a polymer by the activities of the reductase and synthase. The genes encoding these proteins in R. eutropha reside in an operon which has been well characterized (10, 21, 22, 31, 37). Open in a separate windowFIG. 1Pathway for production of PHBV from acetyl-CoA and propionyl-CoA. β-Ketothiolase performs the condensation reactions to generate either acetoacetyl-CoA or β-ketovaleryl-CoA. These are reduced by acetoacetyl-CoA reductase (PhbB) and polymerized by PHB synthase (PhbC).The substrate specificities of these three enzymes are reportedly adequate for production of PHBV copolymer (79), but propionate-fed Escherichia coli harboring the R. eutropha phb operon produces essentially PHB homopolymer (35). Moreover, PHBV copolymer can be produced in E. coli after induction of the fatty acid β-oxidation complex, which contains a β-ketothiolase with broad substrate specificity (26, 27, 35). These data suggest that the R. eutropha PHB pathway is capable of producing copolymer, but only in the context of a second β-ketothiolase with broad substrate specificity.R. eutropha is known to produce at least two β-ketothiolases (7), and at least two distinct plasmid clones which express β-ketothiolase have been isolated from R. eutropha (37). In this work, we analyzed the substrate specificity of the PhbA β-ketothiolase and demonstrated that this enzyme catalyzes thiolysis of β-ketovaleryl-CoA very poorly. We determined that R. eutropha expresses at least two β-ketothiolases in addition to PhbA and that these additional enzymes, which we designate BktB and BktC, efficiently utilize β-ketovaleryl-CoA. We also report the isolation and characterization of bktB (β-ketothiolase B), which encodes the BktB β-ketothiolase required for efficient production of PHBV in R. eutropha.  相似文献   

5.
Recently, a new metabolic link between fatty acid de novo biosynthesis and biosynthesis of poly(3-hydroxy-alkanoate) consisting of medium-chain-length constituents (C6 to C14) (PHAMCL), catalyzed by the 3-hydroxydecanoyl-[acyl-carrier-protein]:CoA transacylase (PhaG), has been identified in Pseudomonas putida (B. H. A. Rehm, N. Krüger, and A. Steinbüchel, J. Biol. Chem. 273:24044–24051, 1998). To establish this PHA-biosynthetic pathway in a non-PHA-accumulating bacterium, we functionally coexpressed phaC1 (encoding PHA synthase 1) from Pseudomonas aeruginosa and phaG (encoding the transacylase) from P. putida in Pseudomonas fragi. The recombinant strains of P. fragi were cultivated on gluconate as the sole carbon source, and PHA accumulation to about 14% of the total cellular dry weight was achieved. The respective polyester was isolated, and GPC analysis revealed a weight average molar mass of about 130,000 g mol−1 and a polydispersity of 2.2. The PHA was composed mainly (60 mol%) of 3-hydroxydecanoate. These data strongly suggested that functional expression of phaC1 and phaG established a new pathway for PHAMCL biosynthesis from nonrelated carbon sources in P. fragi. When fatty acids were used as the carbon source, no PHA accumulation was observed in PHA synthase-expressing P. fragi, whereas application of the β-oxidation inhibitor acrylic acid mediated PHAMCL accumulation. The substrate for the PHA synthase PhaC1 is therefore presumably directly provided through the enzymatic activity of the transacylase PhaG by the conversion of (R)-3-hydroxydecanoyl-ACP to (R)-3-hydroxydecanoyl-CoA when the organism is cultivated on gluconate. Here we demonstrate for the first time the establishment of PHAMCL synthesis from nonrelated carbon sources in a non-PHA-accumulating bacterium, employing fatty acid de novo biosynthesis and the enzymes PhaG (a transacylase) and PhaC1 (a PHA synthase).  相似文献   

6.
(R,S)-[1-14C]3-Hydroxy eicosanoyl-coenzyme A (CoA) has been chemically synthesized to study the 3-hydroxy acyl-CoA dehydratase involved in the acyl-CoA elongase of etiolated leek (Allium porrum L.) seedling microsomes. 3-Hydroxy eicosanoyl-CoA (3-OH C20:0-CoA) dehydration led to the formation of (E)-2,3 eicosanoyl-CoA, which has been characterized. Our kinetic studies have determined the optimal conditions of the dehydration and also resolved the stereospecificity requirement of the dehydratase for (R)-3-OH C20:0-CoA. Isotopic dilution experiments showed that 3-hydroxy acyl-CoA dehydratase had a marked preference for (R)-3-OH C20:0-CoA. Moreover, the very-long-chain synthesis using (R)-3-OH C20:0-CoA isomer and [2-14C]malonyl-CoA was higher than that using the (S) isomer, whatever the malonyl-CoA and the 3-OH C20:0-CoA concentrations. We have also used [1-14C]3-OH C20:0-CoA to investigate the reductant requirement of the enoyl-CoA reductase of the acyl-CoA elongase complex. In the presence of NADPH, [1-14C]3-OH C20:0-CoA conversion was stimulated. Aside from the product of dehydration, i.e. (E)-2,3 eicosanoyl-CoA, we detected eicosanoyl-CoA resulting from the reduction of (E)-2,3 eicosanoyl-CoA. When we replaced NADPH with NADH, the eicosanoyl-CoA was 8- to 10-fold less abundant. Finally, in the presence of malonyl-CoA and NADPH or NADH, [1-14C]3-OH C20:0-CoA led to the synthesis of very-long-chain fatty acids. This synthesis was measured using [1-14C]3-OH C20:0-CoA and malonyl-CoA or (E)-2,3 eicosanoyl-CoA and [2-14C]malonyl-CoA. In both conditions and in the presence of NADPH, the acyl-CoA elongation activity was about 60 nmol mg−1 h−1, which is the highest ever reported for a plant system.  相似文献   

7.
Recombinant Ralstonia eutropha strain PHB4 expressing the broad-substrate-specificity polyhydroxyalkanoate (PHA) synthase 1 from Pseudomonas sp. strain 61-3 (PhaC1Ps) synthesizes a PHA copolymer containing the branched side-chain unit 3-hydroxy-4-methylvalerate (3H4MV), which has a carbon backbone identical to that of leucine. Mutant strain 1F2 was derived from R. eutropha strain PHB4 by chemical mutagenesis and shows higher levels of 3H4MV production than does the parent strain. In this study, to understand the mechanisms underlying the enhanced production of 3H4MV, whole-genome sequencing of strain 1F2 was performed, and the draft genome sequence was compared to that of parent strain PHB4. This analysis uncovered four point mutations in the 1F2 genome. One point mutation was found in the ilvH gene at amino acid position 36 (A36T) of IlvH. ilvH encodes a subunit protein that regulates acetohydroxy acid synthase III (AHAS III). AHAS catalyzes the conversion of pyruvate to 2-acetolactate, which is the first reaction in the biosynthesis of branched amino acids such as leucine and valine. Thus, the A36T IlvH mutation may show AHAS tolerance to feedback inhibition by branched amino acids, thereby increasing carbon flux toward branched amino acid and 3H4MV biosynthesis. Furthermore, a gene dosage study and an isotope tracer study were conducted to investigate the 3H4MV biosynthesis pathway. Based on the observations in these studies, we propose a 3H4MV biosynthesis pathway in R. eutropha that involves a condensation reaction between isobutyryl coenzyme A (isobutyryl-CoA) and acetyl-CoA to form the 3H4MV carbon backbone.  相似文献   

8.
The pathway of propionate conversion in a syntrophic coculture of Smithella propionica and Methanospirillum hungatei JF1 was investigated by 13C-NMR spectroscopy. Cocultures produced acetate and butyrate from propionate. [3-13C]propionate was converted to [2-13C]acetate, with no [1-13C]acetate formed. Butyrate from [3-13C]propionate was labeled at the C2 and C4 positions in a ratio of about 1:1.5. Double-labeled propionate (2,3-13C) yielded not only double-labeled acetate but also single-labeled acetate at the C1 or C2 position. Most butyrate formed from [2,3-13C]propionate was also double labeled in either the C1 and C2 atoms or the C3 and C4 atoms in a ratio of about 1:1.5. Smaller amounts of single-labeled butyrate and other combinations were also produced. 1-13C-labeled propionate yielded both [1-13C]acetate and [2-13C]acetate. When 13C-labeled bicarbonate was present, label was not incorporated into acetate, propionate, or butyrate. In each of the incubations described above, 13C was never recovered in bicarbonate or methane. These results indicate that S. propionica does not degrade propionate via the methyl-malonyl-coenzyme A (CoA) pathway or any other of the known pathways, such as the acryloyl-CoA pathway or the reductive carboxylation pathway. Our results strongly suggest that propionate is dismutated to acetate and butyrate via a six-carbon intermediate.  相似文献   

9.
High precision ultrasonic and densimetric techniques have been used to study the interaction of Ni2+ ions with right-handed poly[d(G-C)]·poly[d(G-C)], poly-[d(A-C)]·poly[d(G-T)] and poly[d(A-T)]·poly[d(A-T)] in 5 mM CsCl, 0.2 mM HEPES, pH 7.5 at 20°C. From these measurements the changes in the apparent molar volume and the apparent molar adiabatic compressibility due to the interaction have been obtained. The volume effects of the binding, calculated per mole of Ni2+ ions, range from 11.7 to 23.9 cm3 mol–1 and the compressibility effects range from 19.3 × 10–4 to 43.1 × 10–4 cm3 mol–1 bar–1. These data are interpreted in terms of dehydration of the polynucleotides and Ni2+ ions, i.e. the release of water molecules from the hydration shells of the molecules. An increase in G+C content gives an increase in volume and compressibility effects, indicating a rise in the extent of dehydration. The dehydration effects of Ni2+ binding to poly[d(G-C)]·poly[d(G-C)] are approximately twice those of poly[d(A-T)]·poly[d(A-T)]. The volume and compressibility effects of Ni2+–EDTA complex formation have also been measured and used as a model system for quantitative estimation. These values revealed that Ni2+ ions can coordinate two atomic groups of poly[d(G-C)]·poly[d(G-C)], while in the case of the Ni2+–poly[d(A-T)]·poly[d(A-T)] complex volume and compressibility effects correspond to one direct or two indirect (through water) contacts.  相似文献   

10.
In previous papers of this series the temperature-dependent Raman spectra of poly(dA)·poly(dT) and poly(dA–dT)·poly(dA–dT) were used to characterize structurally the melting and premelting transitions in DNAs containing consecutive A·T and alternating A·T/T·A base pairs. Here, we describe procedures for obtaining thermodynamic parameters from the Raman data. The method exploits base-specific and backbone-specific Raman markers to determine separate thermodynamic contributions of A, T and deoxyribosyl-phosphate moieties to premelting and melting transitions. Key findings include the following: (i) Both poly(dA)·poly(dT) and poly(dA–dT)· poly(dA–dT) exhibit robust premelting transitions, due predominantly to backbone conformational changes. (ii) The significant van’t Hoff premelting enthalpies of poly(dA)·poly(dT) [ΔHvHpm = 18.0 ± 1.6 kcal·mol–1 (kilocalories per mole cooperative unit)] and poly(dA–dT)·poly(dA–dT) (ΔHvHpm = 13.4 ± 2.5 kcal·mol–1) differ by an amount (~4.6 kcal·mol–1) estimated as the contribution from three-centered inter-base hydrogen bonding in (dA)n·(dT)n tracts. (iii) The overall stacking free energy of poly(dA)· poly(dT) [–6.88 kcal·molbp–1 (kilocalories per mole base pair)] is greater than that of poly(dA–dT)· poly(dA–dT) (–6.31 kcal·molbp–1). (iv) The difference between stacking free energies of A and T is significant in poly(dA)·poly(dT) (ΔΔGst = 0.8 ± 0.3 kcal· molbp–1), but marginal in poly(dA–dT)·poly(dA–dT) (ΔΔGst = 0.3 ± 0.3 kcal·molbp–1). (v) In poly(dA)· poly(dT), the van’t Hoff parameters for melting of A (ΔHvHA = 407 ± 23 kcal·mol–1, ΔSvHA = 1166 ± 67 cal·°K–1·mol–1, ΔGvH(25°C)A = 60.0 ± 3.2 kcal·mol–1) are clearly distinguished from those of T (ΔHvHT = 185 ± 38 kcal·mol–1, ΔSvHT = 516 ± 109 cal·°K–1·mol–1, ΔGvH(25°C)T = 27.1 ± 5.5 kcal·mol–1). (vi) Similar relative differences are observed in poly(dA–dT)· poly(dA–dT) (ΔHvHA = 333 ± 54 kcal·mol–1, ΔSvHA = 961 ± 157 cal·°K–1·mol–1, ΔGvH(25°C)A = 45.0 ± 7.6 kcal· mol–1; ΔHvHT = 213 ± 30 kcal·mol–1, ΔSvHT = 617 ± 86 cal·°K–1·mol–1, ΔGvH(25°C)T = 29.3 ± 4.9 kcal·mol–1). The methodology employed here distinguishes thermodynamic contributions of base stacking, base pairing and backbone conformational ordering in the molecular mechanism of double-helical B DNA formation.  相似文献   

11.
The Streptomyces glaucescens fabH gene, encoding β-ketoacyl-acyl carrier protein (β-ketoacyl-ACP) synthase (KAS) III (FabH), was overexpressed in Escherichia coli, and the resulting gene product was purified to homogeneity by metal chelate chromatography. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis of the purified protein revealed an Mr of 37,000, while gel filtration analysis determined a native Mr of 72,000 ± 3,000 (mean ± standard deviation), indicating that the enzyme is homodimeric. The purified recombinant protein demonstrated both KAS activity and acyl coenzyme A (acyl-CoA):ACP transacylase (ACAT) activity in a 1:0.12 ratio. The KAS and ACAT activities were both sensitive to thiolactomycin inhibition. The KAS activity of the protein demonstrated a Km value of 3.66 μM for the malonyl-ACP substrate and an unusual broad specificity for acyl-CoA substrates, with Km values of 2.4 μM for acetyl-CoA, 0.71 μM for butyryl-CoA, and 0.41 μM for isobutyryl-CoA. These data suggest that the S. glaucescens FabH is responsible for initiating both straight- and branched-chain fatty acid biosynthesis in Streptomyces and that the ratio of the various fatty acids produced by this organism will be dictated by the ratios of the various acyl-CoA substrates that can react with FabH. Results from a series of in vivo directed biosynthetic experiments in which the ratio of these acyl-CoA substrates was varied are consistent with this hypothesis. An additional set of in vivo experiments using thiolactomycin provides support for the role of FabH and further suggests that a FabH-independent pathway for straight-chain fatty acid biosynthesis operates in S. glaucescens.  相似文献   

12.
3-Methyladenine (3MeA) DNA glycosylases initiate base excision repair by removing 3MeA. These glycosylases also remove a broad spectrum of spontaneous and environmentally induced base lesions in vitro. Mouse cells lacking the Aag 3MeA DNA glycosylase (also known as the Mpg, APNG or ANPG DNA glycosylase) are susceptible to 3MeA-induced S phase arrest, chromosome aberrations and apoptosis, but it is not known if Aag is solely responsible for repair of 3MeA in vivo. Here we show that in Aag–/– cells, 3MeA lesions disappear from the genome slightly faster than would be expected by spontaneous depurination alone, suggesting that there may be residual repair of 3MeA. However, repair of 3MeA is at least 10 times slower in Aag–/– cells than in Aag+/+ cells. Consequently, 24 h after exposure to [3H]MNU, 30% of the original 3MeA burden is intact in Aag–/– cells, while 3MeA is undetectable in Aag+/+ cells. Thus, Aag is the major DNA glycosylase for 3MeA repair. We also investigated the in vivo repair kinetics of another Aag substrate, 7-methylguanine. Surprisingly, 7-methylguanine is removed equally efficiently in Aag+/+ and Aag–/– cells, suggesting that another DNA glycosylase acts on lesions previously thought to be repaired by Aag.  相似文献   

13.
Magnesium ions strongly influence the structure and biochemical activity of RNA. The interaction of Mg2+ with an equimolar mixture of poly(rA) and poly(rU) has been investigated by UV spectroscopy, isothermal titration calorimetry, ultrasound velocimetry and densimetry. Measurements in dilute aqueous solutions at 20°C revealed two differ ent processes: (i) Mg2+ binding to unfolded poly(rA)·poly(rU) up to [Mg2+]/[phosphate] = 0.25; and (ii) poly(rA)·2poly(rU) triplex formation at [Mg2+]/[phosphate] between 0.25 and 0.5. The enthalpies of these two different processes are favorable and similar to each other, ~–1.6 kcal mol–1 of base pairs. Volume and compressibility effects of the first process are positive, 8 cm3 mol–1 and 24 × 10–4 cm3 mol–1 bar–1, respectively, and correspond to the release of water molecules from the hydration shells of Mg2+ and the polynucleotides. The triplex formation is also accompanied by a positive change in compressibility, 14 × 10–4 cm3 mol–1 bar–1, but only a small change in volume, 1 cm3 mol–1. A phase diagram has been constructed from the melting experiments of poly(rA)·poly(rU) at a constant K+ concentration, 140 mM, and various amounts of Mg2+. Three discrete regions were observed, corresponding to single-, double- and triple-stranded complexes. The phase boundary corresponding to the transition between double and triple helical conformations lies near physiological salt concentrations and temperature.  相似文献   

14.
15.
We report for the first time the in vitro characterization of a reverse tetracycline repressor (revTetR). The dimeric wild-type repressor (TetR) binds to tet operator tetO in the absence of the inducer anhydrotetracycline (atc) to confer tight repression. We have isolated the revTetR G96E L205S mutant, which, contrary to TetR, binds tetO only in the presence of atc. This reverse acting mutant was overproduced and purified. Effector and DNA binding properties were analyzed by EMSA and quantified by fluorescence titration and surface plasmon resonance. The association constant KA of revTetR for binding of [atcMg]+ is ~108 M–1, four orders of magnitude lower than that of TetR. The affinity of TetR for tetO is 5.6 ± 2 × 109 M–1 and that for revTetR in the presence of atc is 1 ± 0.2 × 108 M–1. Both induced forms, the atc-bound TetR and the free revTetR, have the same low affinity of 4 ± 1 × 105 M–1 for DNA. Therefore, atc does not act as a dimerization agent for revTetR. We discuss the structural differences between TetR and revTetR potentially underlying this reversal of activity.  相似文献   

16.
[URE3] is a prion of the yeast Ure2 protein. Hsp40 is a cochaperone that regulates Hsp70 chaperone activity. When overexpressed, the Hsp40 Ydj1p cures yeast of [URE3], but the Hsp40 Sis1p does not. On the basis of biochemical data Ydj1p has been proposed to cure [URE3] by binding soluble Ure2p and preventing it from joining prion aggregates. Here, we mutagenized Ydj1p and find that disrupting substrate binding, dimerization, membrane association, or ability to transfer substrate to Hsp70 had little or no effect on curing. J-domain point mutations that disrupt functional interactions of Ydj1p with Hsp70 abolished curing, and the J domain alone cured [URE3]. Consistent with heterologous J domains possessing similar Hsp70 regulatory activity, the Sis1p J domain also cured [URE3]. We further show that Ydj1p is not essential for [URE3] propagation and that depletion of Ure2p is lethal in cells lacking Ydj1p. Our data imply that curing of [URE3] by overproduced Ydj1p does not involve direct interaction of Ydj1p with Ure2p but rather works through regulation of Hsp70 through a specific J-protein/Hsp70 interaction.  相似文献   

17.
Mycobacterium tuberculosis H37Ra and M. bovis BCG produce multiple methyl-branched fatty acids called mycocerosic acids, presumably from methyl-malonyl coenzyme A (CoA). An acyl-CoA carboxylase was isolated from these organisms at a 30 to 50% yield by a purification procedure involving ammonium sulfate fractionation, gel filtration, and affinity chromatography with a monomeric avidin–Sepharose 4B-CL gel with d-biotin as the eluant. Sodium dodecyl sulfate electrophoresis and avidin binding indicate that each enzyme is probably composed of two dissimilar subunits with a covalently bound biotin in the larger subunit. The enzyme preparations from H37Ra and BCG had specific activities of 2.1 and 5.5 μmol min−1 mg−1, respectively, when propionyl-CoA was the substrate. The enzymes from the two species displayed striking similarities in their kinetic parameters. They showed maximal activity at pH 8.0 when propionyl-CoA was the substrate, but displayed a relatively broad pH-activity profile when acetyl-CoA was the substrate. With both substrates, potassium phosphate buffer gave maximal activity. Apparent Km values for propionyl-CoA, ATP, Mg2+, and NaHCO3 were 70 μM, 100 μM, 5.4 mM, and 2.2 mM, respectively. The enzyme also carboxylated acetyl-CoA and butyryl-CoA, and high-performance liquid chromatography showed the expected products of carboxylation. However, with these substrates, the Km was higher and the Vmax was lower than those of propionyl-CoA. The enzyme was shown to be stereospecific, synthesizing exclusively (S)-methylmalonyl-CoA from propionyl-CoA. No other acyl-CoA carboxylase was observed during the purification procedure, indicating that the present carboxylase may provide malonyl-CoA for the synthesis of n-fatty acids as well as methylmalonyl-CoA for the synthesis of mycocerosic acids.  相似文献   

18.
1. The rate and stability to aging of the metabolism of propionate by sheep-liver slices and sucrose homogenates were examined. Aging for up to 20min. at 37° in the absence of added substrate had little effect with slices, whole homogenates or homogenates without the nuclear fraction. 2. Metabolism of propionate by sucrose homogenates was confined to the mitochondrial fraction, but the mitochondrial supernatant (microsomes plus cell sap) stimulated propionate removal. 3. The rate of propionate metabolism by liver slices was higher in a high potassium phosphate–bicarbonate medium [0·88(±s.e.m. 0·16)μmole/mg. of N/hr.] than in Krebs–Ringer bicarbonate medium [0·44(±s.e.m. 0·13)μmole/mg. of N/hr.]. 4. Metabolism of propionate by sucrose homogenates freed from nuclei was dependent on the presence of oxygen, carbon dioxide and ATP. Propionate removal was stimulated 250% by Mg2+ ions and 670% by cytochrome c. 5. In the complete medium 2·39(±s.e.m. 0·15)μmoles of propionate were consumed/mg. of N/hr. 6. The ratio of oxygen consumption to propionate utilization was sufficient to account for the complete oxidation of half the propionate consumed. 7. The only products detected under these conditions were succinate, fumarate and malate. Propionate had no effect on the production of lactate from endogenous sources and did not itself give rise to lactate. 8. Methylmalonate did not accumulate when propionate was metabolized and was not oxidized. It was detected as an intermediate in the conversion of propionyl-CoA into succinate. The rate of this reaction sequence was adequate to account for the rate of propionate metabolism by sucrose homogenates or slices, provided that the rate of formation of propionyl-CoA was not limiting. 9. The methylmalonate pathway was predominantly a mitochondrial function. 10. The metabolism of propionate appeared to be dependent on active oxidative phosphorylation.  相似文献   

19.
1. Rates of entry and oxidation of a range of metabolites have been measured in tracheostomized sheep (diet, 800g. of lucerne chaff and 100g. of maize/day) by combining isotope-dilution techniques with the continuous measurement of total respiratory gas exchange, and 14CO2 production during the intravenous or intraruminal infusion of 14C-labelled substrates. 2. Mean entry rates in fed and starved (24hr.) sheep respectively, expressed as mg./min./kg. body wt.0·75, were: glucose, 5·0 (range 4·8–5·1, 2 observations) and 3·8 (3·2–4·2, 4); acetate, 10·8 (9·1–13·5, 4) and 5·8 (1); d(−)-β-hydroxybutyrate, 1·4 (1) and 1·5 (0·8–2·4, 4); palmitate, oleate and stearate (starved sheep only) 1·0 (0·6–1·9, 7), 0·9 (0·2–1·6, 10) and 0·9 (0·5–1·1, 11) respectively. 3. Production rates of propionate and butyrate in continuously feeding sheep were 6·4 (4·7–8·3, 4) and 4·3 (3·4–6·1, 4) mg./min./kg.0·75 respectively, and in starved (24hr.) sheep were 2·5 (2·2–2·9, 2) and 1·0 (0·8–1·2, 2) mg./min./kg.0·75 respectively. 4. Calculated terminal values for the specific radioactivity of respiratory 14CO2 during measurements of entry rates and production rates were used to calculate the contributions of individual substrates to overall oxidative metabolism. Mean values for fed and starved sheep respectively were: glucose, 9·1 (8·6–9·6, 2) and 11·2 (5·9–15·1, 4)%; acetate, 31·6 (26·8–38·1, 4) and 22·1 (1)%; d(−)-β-hydroxybutyrate, 10·4 (1) and 4·8 (1·9–7·7, 4)%; propionate, 23·0 (13·8–29·9, 4) and 7·1 (6·8–7·4, 2)%; butyrate, 16·5 (13·7–20·5, 4) and 5·3 (5·2–5·3, 2)%; palmitate, oleate and stearate (starved sheep only), 4·7 (2·0–7·7, 7), 4·0 (1·2–6·6, 10) and 4·4 (3·8–5·8, 9)% respectively. The sum of these values for individual substrates in fed and starved sheep, excluding that of β-hydroxybutyrate and after correction of the glucose value for the known interrelations of this substrate with propionate, accounted for 76% and 58% respectively of total production of carbon dioxide. 5. Calculations based on the proportion of substrate entry directly oxidized indicated that the substrates studied accounted for 63% (fed sheep) and 43% (starved sheep) of total energy expenditure measured by oxygen uptake. The contribution of β-hydroxybutyrate was excluded, and corrections were made for glucose–propionate interrelations, and for the different rates of oxidation of the methyl and carboxyl fragments of acetate. 6. The present results have been combined with those obtained earlier in this Laboratory to examine the relationships between rates of substrate entry and oxidation, and concentrations of substrate in blood. Rates of entry of acetate, glucose, d(−)-β-hydroxybutyrate, palmitate and oleate (but not stearate) were well correlated with concentration in blood, and substrate contribution to production of carbon dioxide showed a similar correlation to blood concentration, except with glucose. 7. It was concluded that the general technique is of potential value in providing valid quantitative parameters of animal metabolism.  相似文献   

20.
We developed an isotopic technique to assess mitochondrial acetyl-CoA turnover (≈citric acid flux) in perfused rat hearts. Hearts are perfused with buffer containing tracer [13C2,2H3]acetate, which forms M5 + M4 + M3 acetyl-CoA. The buffer may also contain one or two labeled substrates, which generate M2 acetyl-CoA (e.g. [13C6]glucose or [1,2-13C2]palmitate) or/and M1 acetyl-CoA (e.g. [1-13C]octanoate). The total acetyl-CoA turnover and the contributions of fuels to acetyl-CoA are calculated from the uptake of the acetate tracer and the mass isotopomer distribution of acetyl-CoA. The method was applied to measurements of acetyl-CoA turnover under different conditions (glucose ± palmitate ± insulin ± dichloroacetate). The data revealed (i) substrate cycling between glycogen and glucose-6-P and between glucose-6-P and triose phosphates, (ii) the release of small excess acetyl groups as acetylcarnitine and ketone bodies, and (iii) the channeling of mitochondrial acetyl-CoA from pyruvate dehydrogenase to carnitine acetyltransferase. Because of this channeling, the labeling of acetylcarnitine and ketone bodies released by the heart are not proxies of the labeling of mitochondrial acetyl-CoA.  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