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.
Uptake and degradation of EDTA by Escherichia coli   总被引:1,自引:0,他引:1  
It was found that Escherichia coli exhibited a growth by utilization of Fe(III)EDTA as a sole nitrogen source. No significant growth was detected when Fe(III)EDTA was replaced by EDTA complexes with other metal ions such as Ca2+, Co2+, Cu2+, Mg2+, Mn2+, and Zn2+. When EDTA uptake was measured in the presence of various ions, it was remarkable only when Fe3+ was present. The cell extract of E. coli exhibited a significant degradation of EDTA only in the presence of Fe3+. It is likely that the capability of E. coli for the growth by utilization of Fe(III)EDTA results from the Fe3+-dependent uptake and degradation of EDTA.  相似文献   

2.
Glu-69 belongs to a proposed active-site consensus motif His62-X-His64-X4-Glu69 (where X is any amino acid) that acetylacetone dioxygenase Dke1 from Acinetobacter johnsonii shares with structurally related non-heme metal enzymes of the cupin protein superfamily. We report functional consequences of the site-directed replacement Glu-69 → Gln based on a detailed biochemical and kinetic characterization of the purified Dke1 mutant. Perturbations of the free energy profile of the wild-type caused by the mutation were surprisingly small, with key points of the reaction pathway such as β-diketone substrate binding, the rate-limiting reduction of dioxygen, and CC bond cleavage essentially left unaltered. Release of Fe2+ from the mutant active site occurred at twice the wild-type rate, and the thermal stability of β-sheet secondary structure in Fe2+-depleted apo-proteins was lower in the mutant. The substitution Glu-69 → Gln is thus remarkably silent regarding Dke1 function. These results do not support a unified catalytic or metal-coordinating role of Glu-69 (and its positional homologues) in O2-dependent cupin-fold enzymes.  相似文献   

3.
The main aim of this study was the determination of the operational stability of soluble Dke1 (EC 1.13.11.50) in an enzyme membrane reactor. In order to calculate the half-life of soluble Dke1, the K M of oxygen must be known. The determination of this constant was done using progress curve analysis (K M=260 μmol l−1). In a next step, the reactor system was studied by building a mathematical model for calculation of the reactor system, using Berkeley Madonna ver. 8.0.1 software. After that, the determination of the half-life of Dke1 under operational conditions at different temperatures (5, 10, 15, 25, 30, 35°C) was performed. The quantitative criterion for stability was the value of the first-order rate constant of monomolecular inactivation. The experiments showed that soluble Dke1 is poorly stable. The half-life ranged from 308 min at 5°C to 9 min at 35°C. This method for determining the half-life is quite applicable for enzymes which are poorly stable. In addition, both the storage stability and the operational stability can be determined.  相似文献   

4.
Previously, we reported that pyruvate production was markedly improved in TBLA-1, an H+-ATPase-defective Escherichia coli mutant derived from W1485lip2, a pyruvate-producing E. coli K-12 strain. TBLA-1 produced more than 30 g/l pyruvate from 50 g/l glucose by jar fermentation, while W1485lip2 produced only 25 g/l pyruvate (Yokota et al. in Biosci Biotechnol Biochem 58:2164–2167, 1994b). In this study, we tested the ability of TBLA-1 to produce alanine by fermentation. The alanine dehydrogenase (ADH) gene from Bacillus stearothermophilus was introduced into TBLA-1, and direct fermentation of alanine from glucose was carried out. However, a considerable amount of lactate was also produced. To reduce lactate accumulation, we knocked out the lactate dehydrogenase gene (ldhA) in TBLA-1. This alanine dehydrogenase-expressing and lactate dehydrogenase-defective mutant of TBLA-1 produced 20 g/l alanine from 50 g/l glucose after 24 h of fermentation. The molar conversion ratio of glucose to alanine was 41%, which is the highest level of alanine production reported to date. This is the first report to show that an H+-ATPase-defective mutant of E. coli can be used for amino acid production. Our results further indicate that H+-ATPase-defective mutants may be used for fermentative production of various compounds, including alanine.  相似文献   

5.
Streptococcus thermophilus (ST) chromosomal DNA (chr DNA) fragments having promoter activity were cloned and selected inEscherichia coli using a chloramphenicol acetyltransferase- (cat-) based promoter-probe vector pKK520-3. Insertion of a promoterless streptomycete melanin biosynthesis operon (melC) downstream from the promoters of the library further identified clone STP2201 as a strong promoter inE. coli. Subcloning of a STP2201-melC DNA fragment into the pMEU-seriesS. thermophilus-E. coli shuttle vectors yielded pEU5xML2201x plasmids that conferred Mel+ phenotype toE. coli. The pEU5aML2201a was further shown to afford a high level of tyrosinase production (2 units mg–1 protein) inE. coli, and to produce an apparently inactivemelC gene product that reacts with anti-tyrosinase antiserum inS. thermophilus. SubstitutingmelC with a streptomycete cholesterol oxidase gene (choA) in the same orientation yielded pEU5aCH2201a that conferred ChoA activity to anE. coli transformant at a level of (1.06±0.15)×10–7 units mg–1 protein. Introduction of this plasmid intoS. thermophilus by electrotransformation yielded ChoA transformant that produced the enzyme at about 25% of the level found inE. coli.  相似文献   

6.
Zymomonas mobilis phoA gene encoding alkaline phosphatase was expressed inEscherichia coli CC118 carrying the recombinant plasmid pZAP1. The pH optimum for this enzyme was 9.0 and showed a peak activity at 42°C. This enzyme required Zn2+ for its catalytic activity; however, Mg2+ or Ca2+ significantly affected the activity. This enzyme was found to be ethanolabile, and ethanol inhibition was reversed by addition of Zn2+. Kinetics ofZ. mobilis alkaline phosphatase production inE. coli CC118 (pZAP1) showed that the enzyme activity was growth associated and localized in the cellular fraction, and the maximum activity was found in the stationary phase.  相似文献   

7.
A labeling scheme is introduced that facilitates the measurement of accurate 13Cβ chemical shifts of invisible, excited states of proteins by relaxation dispersion NMR spectroscopy. The approach makes use of protein over-expression in a strain of E. coli in which the TCA cycle enzyme succinate dehydrogenase is knocked out, leading to the production of samples with high levels of 13C enrichment (30–40%) at Cβ side-chain carbon positions for 15 of the amino acids with little 13C label at positions one bond removed (≈5%). A pair of samples are produced using [1-13C]-glucose/NaH12CO3 or [2-13C]-glucose as carbon sources with isolated and enriched (>30%) 13Cβ positions for 11 and 4 residues, respectively. The efficacy of the labeling procedure is established by NMR spectroscopy. The utility of such samples for measurement of 13Cβ chemical shifts of invisible, excited states in exchange with visible, ground conformations is confirmed by relaxation dispersion studies of a protein–ligand binding exchange reaction in which the extracted chemical shift differences from dispersion profiles compare favorably with those obtained directly from measurements on ligand free and fully bound protein samples.  相似文献   

8.
l-Lactate-driven ferric and nitrate reduction was studied in Escherichia coli E4. Ferric iron reduction activity in E. coli E4 was found to be constitutive. Contrary to nitrate, ferric iron could not be used as electron acceptor for growth. Ferric iron reductase activity of 9 nmol Fe2+ mg-1 protein min-1 could not be inhibited by inhibitors for the respiratory chain, like Rotenone, quinacrine, Actinomycin A, or potassium cyanide. Active cells and l-lactate-driven nitrate respiration in E. coli E4 leading to the production of nitrite, was reduced to about 20% of its maximum activity with 5 mM ferric iron, or to about 50% in presence of 5 mM ferrous iron. The inhibition was caused by nitric oxide formed by a purely chemical reduction of nitrite by ferrous iron. Nitric oxide was further chemically reduced by ferrous iron to nitrous oxide. With electron paramagnetic resonance spectroscopy, the presence of a free [Fe2+-NO] complex was shown. In presence of ferrous or ferric iron and l-lactate, nitrate was anaerobically converted to nitric oxide and nitrous oxide by the combined action of E. coli E4 and chemical reduction reactions (chemodenitrification).  相似文献   

9.
Treatment of E. coli extract with iron/ascorbate preferentially inactivated NADP-isocitrate dehydrogenase without affecting glucose-6-phosphate dehydrogenase. NADP-Isocitrate dehydrogenase required divalent metals such as Mg2+, Mn2+ or Fe2+ ion. Iron/ascorbate-dependent inactivation of the enzyme was accompanied with the protein fragmentation as judged by SDS-PAGE. Catalase protecting the enzyme from the inactivation suggests that hydroxyl radical is responsible for the inactivation with fragmentation. TOF-MS analysis showed that molecular masses of the enzyme fragments were 36 and 12, and 33 and 14 kDa as minor components. Based on the amino acid sequence analyses of the fragments, cleavage sites of the enzyme were identified as Asp307-Tyr308 and Ala282-Asp283, which are presumed to be the metal-binding sites. Ferrous ion bound to the metal-binding sites of the E. coli NADP-isocitrate dehydrogenase may generate superoxide radical that forms hydrogen peroxide and further hydroxyl radical, causing inactivation with peptide cleavage of the enzyme. Oxidative inactivation of NADP-isocitrate dehydrogenase without affecting glucose 6-phosphate dehydrogenase shows only a little influence on the antioxidant activity supplying NADPH for glutathione regeneration, but may facilitate flux through the glyoxylate bypass as the biosynthetic pathway with the inhibition of the citric acid cycle under aerobic growth conditions of E. coli.  相似文献   

10.
A novel esterase gene (e25) was identified from Altererythrobacter epoxidivorans CGMCC 1.7731T by genome sequence screening. The e25 gene is 948 nucleotides in length and encodes a 315?amino acid protein (E25) with a predicted molecular mass of 33,683 Da. A phylogenetic tree revealed that E25 belongs to the hormone-sensitive lipase (HSL) family of lipolytic enzymes. An activity assay of E25 showed that it exhibited the highest catalytic efficiency when using p-nitrophenyl caproate (C6) as a substrate. The optimum pH and temperature were determined to be approximately pH 9 and 45°C, and the Km and Vmax values were 0.12?mM and 1,772?µmol/min/mg, respectively. After an incubation at 40°C for 80?min, E25 retained 75% of its basal activity. The enzyme exhibited good tolerance to metal cations, such as Ba2+, Ca2+, and Cu2+ (10?mM), but its activity was strongly inhibited by Co2+, Ni2+, Mn2+, and Zn2+. The E25 enzyme was stimulated by glycerol and retained over 60% of its basal activity in the presence of 1% Tween-80 and Triton X-100. Overall, the activity of E25 under alkaline conditions and its organic solvent and detergent tolerance indicate that E25 could be useful as a novel industrial catalyst in biotechnological applications.  相似文献   

11.
The pathway leading to the formation of ethylene as a secondary metabolite from methionine by Escherichia coli strain B SPAO has been investigated. Methionine was converted to 2-oxo-4-methylthiobutyric acid (KMBA) by a soluble transaminase enzyme. 2-Hydroxy-4-methylthiobutyric acid (HMBA) was also a product, but is probably not an intermediate in the ethylene-forming pathway. KMBA was converted to ethylene, methanethiol and probably carbon dioxide by a soluble enzyme system requiring the presence of NAD(P)H, Fe3+ chelated to EDTA, and oxygen. In the absence of added NAD(P)H, ethylene formation by cell-free extracts from KMBA was stimulated by glucose. The transaminase enzyme may allow the amino group to be salvaged from methionine as a source of nitrogen for growth. As in the plant system, ethylene produced by E. coli was derived from the C-3 and C-4 atoms of methionine, but the pathway of formation was different. It seems possible that ethylene production by bacteria might generally occur via the route seen in E. coli.Abbreviations EDTA ethylenediaminetetraacetic acid - HMBA 2-hydroxy-4-methylthiobutyric acid (methionine hydroxy analogue) - HSS high speed supernatant - KMBA 2-oxo-4-methylthiobutyric acid - PCS phase combining system  相似文献   

12.
 Four reductions of the R2 subunit of mouse ribonucleotide reductase have been studied and found to exhibit different behaviour from that of Escherichia coli R2. An important difference is that there is no stable met-R2 (Fe2 II I) form of mouse R2. With hydroxyurea, hydrazine and hydroxylamine uniphasic kinetics are observed for the combined reduction of radical Tyr ˙ and Fe2 II I components to tyrosine and Fe2 II respectively. The rate constants, determined at 370 nm (emphasising FeIII decay) and 417 nm (emphasising Tyr ˙ decay), differ by factors of 2–3, allowing some mechanistic features to be defined. The studies with hydrazine are particularly important. In the case of E. coli R2, a first phase corresponding to two-equivalent reduction of the met-R2 component has been observed [18]. It is likely that the four times slower second phase reaction of active E. coli R2 also corresponds to the Fe2 II I → Fe2 II change and is followed by fast intramolecular Fe2 II reduction of the higher potential Tyr ˙. The latter changes are believed to hold also for (active) mouse R2. The FeIIFeIII semi forms have been detected at low levels by EPR for mouse R2 (9%) and E. coli (∼5%) in previous studies. Further substrate reduction of FeIIFeIII occurs at a comparable rate to account for the transient behaviour of FeIIFeIII. For mouse R2 the combined FeIII decay processes (which we are unable to separate) give smaller uniphasic rate constants at 370 nm than at 417 nm. A fitted-base-line (FBL) treatment of absorbance changes at 417 nm targets more closely the Tyr ˙ decay as a means of monitoring the rate-determining step. The FBL method gives rate constants k (M–1 s–1) at 25  °C and pH 7.5 for hydroxyurea (1.46), hydrazine (0.163) and hydroxylamine (4.4). Surprisingly, phenylhydrazine, with a less strong reduction potential (0.25 V), gives a substantially faster reduction of the Tyr ˙ as the only redox step (rate constant 27 M–1 s–1). In this case a slower second phase at 370 nm is independent of reductant and corresponds to rate-controlling release of FeIII. Overall the results indicate a more reactive redox centre for mouse R2 and help develop further an understanding of factors affecting the reactivity of R2. Received: 11 October 1996 / Accepted: 11 February 1997  相似文献   

13.
Bacillus sp. GRE1 isolated from an Ethiopian hyperthermal spring produced raw-starch digesting, Ca2+-independent thermostable α-amylase. Enzyme production in shake flask experiments using optimum nutrient supplements and environmental conditions was 2,360 U l−1. Gel filtration chromatography yielded a purification factor of 33.6-fold and a recovery of 46.5%. The apparent molecular weight of the enzyme was 55 kDa as determined by SDS-PAGE. Presence or absence of Ca2+ produced similar temperature optima of 65–70°C. The optimum pH was in the range of 5.5–6.0. The enzyme maintained 50% of its original activity after 45 min of incubation at 80°C and was stable at a pH range of 5.0–9.0. The V max and K m values for soluble starch were 42 mg reducing sugar min−1 and 4.98 mg starch ml−1, respectively. Strong inhibitors of enzyme activity included Cu2+, Zn2+ and Fe2+. The enzyme coding gene and the deduced protein translation revealed a characteristic but markedly atypical homology to Bacillus species α-amylase sequences. The enzyme hydrolyzed wheat, corn and tapioca starch granules efficiently below their gelatinization temperatures. Rather than the higher oligosaccharides normally produced by Bacillus α-amylases operating at high temperatures, maltose was the major hydrolysis product with the present enzyme.  相似文献   

14.
The cellobiohydrolase II (CBHII) of Microbispora bispora, originally cloned in Escherichia coli, was subcloned into Streptomyces lividans using shuttle vectors pSKN 01 and pSKN 02. The enzyme was secreted from Streptomyces, whereas it was intracellular in E. coli. The yields of CBHII produced by S. lividans transformants were 15–20-fold higher than those produced by E. coli transformants. The optimal pH of M. bispora native cellobiohydrolase and the cloned enzyme from S. lividans is 6.5. The thermal and pH stability of CHBII produced in M. bispora, E. coli and S. lividans were compared. Enzyme produced in E. coli was inactivated more rapidly (k = 0.252 min–1 at 90° C; 90% inactivation after 10 min vs. 0.119 min–1 for the others). CBHII was monitored following electrophoretic separation by reaction with a monoclonal antibody. The apparent molecular mass of the protein produced from the S. lividans clone was 93 kDa, the same as that of the native enzyme, but that of the enzyme produced in E. coli was smaller (82 kDa). Correspondence to: P. Hu  相似文献   

15.
In the industrial processing of starch for sugar syrup and ethanol production, a liquefaction step is involved where starch is initially solubilized at high temperature and partially hydrolyzed with a thermostable and thermoactive α‐amylase. Most amylases require calcium as a cofactor for their activity and stability, therefore calcium, along with the thermostable enzyme, are typically added to the starch mixture during enzymatic liquefaction, thereby increasing process costs. An attractive alternative would be to produce the enzyme directly in the tissue to be treated. In a proof of concept study, tobacco cell cultures were used as model system to test in planta production of a hyperthermophilic α‐amylase from Thermotoga maritima. While comparable biochemical properties to recombinant production in Escherichia coli were observed, thermostability of the plant‐produced α‐amylase benefited significantly from high intrinsic calcium levels in the tobacco cells. The plant‐made enzyme retained 85% of its initial activity after 3 h incubation at 100°C, whereas the E. coli‐produced enzyme was completely inactivated after 30 min under the same conditions. The addition of Ca2+ or plant cell extracts from tobacco and sweetpotato to the E. coli‐produced enzyme resulted in a similar stabilization, demonstrating the importance of a calcium‐rich environment for thermostability, as well as the advantage of producing this enzyme directly in plant cells where calcium is readily available. Biotechnol. Bioeng. 2009; 104: 947–956. © 2009 Wiley Periodicals, Inc.  相似文献   

16.
Escherichia coli K1 invasion of microvascular endothelial cells of human brain (HBMEC) is required for E. coli penetration into the central nervous system, but the microbial-host interactions that are involved in this invasion of HBMEC remain incompletely understood. We have previously shown that FimH, one of the E. coli determinants contributing to the binding to and invasion of HBMEC, induces Ca2+ changes in HBMEC. In the present study, we have investigated in detail the role of cellular calcium signaling in the E. coli K1 invasion of HBMEC, the main constituents of the blood-brain barrier. Addition of the meningitis-causing E. coli K1 strain RS218 (O18:K1) to HBMEC results in transient increases of intracellular free Ca2+. Inhibition of phospholipase C with U-73122 and the chelating of intracellular Ca2+ by BAPTA/AM reduces bacterial invasion of HBMEC by approximately 50%. Blocking of transmembrane Ca2+ fluxes by extracellular lanthanum ions also inhibits the E. coli invasion of HBMEC by approximately 50%. In addition, E. coli K1 invasion is significantly inhibited when HBMEC are pretreated by the calmodulin antagonists, trifluoperazine or calmidazolium, or by ML-7, a specific inhibitor of Ca2+/calmodulin-dependent myosin light-chain kinase. These findings indicate that host intracellular Ca2+ signaling contributes in part to E. coli K1 invasion of HBMEC. This work was supported by the American Heart Association (grant SDG 0435177N to Y.K.) and by NIH grants (to K.S.K.).  相似文献   

17.
The interaction of outer membrane protein A (OmpA) with its receptor, Ecgp96 (a homologue of Hsp90β), is critical for the pathogenesis of Escherichia coli K1 meningitis. Since Hsp90 chaperones Toll‐like receptors (TLRs), we examined the role of TLRs in E. coli K1 infection. Herein, we show that newborn TLR2?/? mice are resistant to E. coli K1 meningitis, while TLR4?/? mice succumb to infection sooner. In vitro, OmpA+ E. coli infection selectively upregulates Ecgp96 and TLR2 in human brain microvascular endothelial cells (HBMEC), whereas OmpA? E. coli upregulates TLR4 in these cells. Furthermore, infection with OmpA+ E. coli causes Ecgp96 and TLR2 translocate to the plasma membrane of HBMEC as a complex. Immunoprecipitation studies of the plasma membrane fractions from infected HBMEC reveal that the C termini of Ecgp96 and TLR2 are critical for OmpA+ E. coli invasion. Knockdown of TLR2 using siRNA results in inefficient membrane translocation of Ecgp96 and significantly reduces invasion. In addition, the interaction of Ecgp96 andTLR2 induces a bipartite signal, one from Ecgp96 through PKC‐α while the other from TLR2 through MyD88, ERK1/2 and NF‐κB. This bipartite signal ultimately culminates in the efficient production of NO, which in turn promotes E. coli K1 invasion of HBMEC.  相似文献   

18.
Summary Escherichia coli Rl is an Ag+-resistant strain that, as we have shown recently, harbours at least two large plasmids, pJT1 (83 kb) and pJT2 (77 kb). Tn5-Mob was introduced into theE. coli Rl host replicon via conjugation on membrane filters. The transfer functions of plasmid RP4-4 were provided in this process and Tn5-Mob clones mated withE. coli C600 yielded Ag+-resistant transconjugants. This mobilization procedure allowed transfer and expression of pJT1 Ag+ resistance inE. coli C600. Prior to use of Tn5-Mob mobilization, it was not possible to transfer Ag+-resistant determinant(s) intoE. coli by conjugation or transformation including high-voltage electroporation.E. coli C600 containing PJTI and PJT2 displayed decreased accumulation of Ag+ similar toE. coli R1.E. coli C600 could not tolerate 0.1 and 0.5 mM Ag+, rapidly accumulated Ag+ and became non-viable. Tn5-Mob mobilization may be useful in the study of metal resistance in bacteria, especially in strains not studied for resistance mechanisms.  相似文献   

19.
The gene coding for the glutaryl 7-aminocephalosporanic acid (GL 7-ACA) acylase from Pseudomonas diminuta KAC-1 was cloned and expressed in Escherichia coli. The acylase gene was composed of 2160 base pairs and encoded a polypeptide of 720 amino acid residues. The E. coli BL21 carrying pET2, the plasmid construct for high expression of GL 7-ACA acylase gene, produced this enzyme at approx. 30% of the total proteins with 3.2 units activity mg protein–1. Growth at temperature below 31 °C and deletion of signal peptide increased the processing of precursor acylase to active enzyme in the recombinant E. coli cells.  相似文献   

20.
Previous kinetic characterization of Escherichia coli fructose 1,6-bisphosphatase (FBPase) was performed on enzyme with an estimated purity of only 50%. Contradictory kinetic properties of the partially purified E. coli FBPase have been reported in regard to AMP cooperativity and inactivation by fructose-2,6-bisphosphate. In this investigation, a new purification for E. coli FBPase has been devised yielding enzyme with purity levels as high as 98%. This highly purified E. coli FBPase was characterized and the data compared to that for the pig kidney enzyme. Also, a homology model was created based upon the known three-dimensional structure of the pig kidney enzyme. The kcat of the E. coli FBPase was 14.6 s−1 as compared to 21 s−1 for the pig kidney enzyme, while the Km of the E. coli enzyme was approximately 10-fold higher than that of the pig kidney enzyme. The concentration of Mg2+ required to bring E. coli FBPase to half maximal activity was estimated to be 0.62 mM Mg2+, which is twice that required for the pig kidney enzyme. Unlike the pig kidney enzyme, the Mg2+ activation of the E. coli FBPase is not cooperative. AMP inhibition of mammalian FBPases is cooperative with a Hill coefficient of 2; however, the E. coli FBPase displays no cooperativity. Although cooperativity is not observed, the E. coli and pig kidney enzymes show similar AMP affinity. The quaternary structure of the E. coli enzyme is tetrameric, although higher molecular mass aggregates were also observed. The homology model of the E. coli enzyme indicated slight variations in the ligand-binding pockets compared to the pig kidney enzyme. The homology model of the E. coli enzyme also identified significant changes in the interfaces between the subunits, indicating possible changes in the path of communication of the allosteric signal.  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