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 actinomycete Amycolatopsis japonicum is the producer of the chelating compound [S,S]-ethylenediamine-disuccinc acid (EDDS). [S,S]-EDDS is an isomer of ethylenediamine-tetraacetic acid (EDTA), an economically important chelating compound that suffers from an extremely poor degradability. Frequent use of the persistent EDTA in various industrial and domestic applications has caused an accumulation of EDTA in soil as well as in aqueous environments. As a consequence, EDTA is the highest concentrated anthropogenic compound present in water reservoirs. The [S,S]-form of EDDS has chelating properties similar to EDTA, however, in contrast to EDTA it is readily biodegradable. In order to compete with the cost-effective chemical synthesis of EDTA, we aimed to optimize the biotechnological production of [S,S]-EDDS in A. japonicum by using metabolic engineering approaches. Firstly, we integrated several copies of the [S,S]-EDDS biosynthetic genes into the chromosome of A. japonicum and replaced the native zinc responsive promoter with the strong synthetic constitutive promoter SP44*. Secondly, we increased the supply of O-phospho-serine, the direct precursor of [S,S]-EDDS. The combination of these approaches together with the optimized fermentation process led to a significant improvement in [S,S]-EDDS up to 9.8 g/L with a production rate of 4.3 mg/h/g DCW.  相似文献   

2.
Effect of high boron application on boron content and growth of melons   总被引:4,自引:0,他引:4  
Synthetic chelates, such as ethylene diamine tetraacetic acid (EDTA), have been shown to enhance phytoextraction of Pb from contaminated soil but also cause leaching of heavy metal-chelate complexes, posing a groundwater contamination threat. In a soil column study, we examined the effect of EDTA and a biodegradable chelate [S,S] isomere of ethylene diamine disuccinate ([S,S]-EDDS), newly introduced in phytoextraction research, on the uptake of Pb by the Chinese cabbage (Brassica rapa) and Pb leaching through the soil profile. Soil water sorption characteristics were modified by acrylamide hydrogel. The addition of 0.1 and 0.2% (w/w) of hydrogel amendments increased soil field water capacity from initial 24.6% to 28.5% and 31.3%, respectively. The additions of 2.5, 5 and 10 mmol EDTA kg–1 soil were more effective in enhancing Pb plant uptake than comparable [S,S]-EDDS treatments, but caused (as also 10 mmol kg–1 [S,S]-EDDS additions) unacceptably high Pb leaching in treatments with any soil water sorption conditions tested. The most efficient level of EDTA (10 mmol kg–1) enhanced plant Pb uptake by 97 times compared to the control. Shoots Pb concentrations reached 500 mg kg–1 of dry biomass. However, in this treatment 36.2% of total initial Pb was leached from the soil during the first four weeks after chelate addition. Hydrogel soil amendments were more effective in treatments with [S,S]-EDDS than with EDTA. In treatments with 10 mmol kg–1[S,S]-EDDS hydrogel amended soils, plant Pb uptake was significantly reduced and Pb leach was as high as 44.2% of total initial soil Pb. At lower [S,S]-EDDS concentrations, the effect of hydrogel soil amendment on Pb leaching was the opposite. The addition of 5 mmol kg–1 [S,S]-EDDS soil to the soil amended with 0.2% hydrogel increased Pb uptake by 18 times while only 0.2% of total initial Pb was leached. In all treatments, the concentrations of Pb in dry plant biomass were far from concentrations required for efficient soil remediation within a reasonable time span.  相似文献   

3.
Abstract

Column leaching using [S, S]-ethylene diamine disuccinate ([S, S]-EDDS) on copper tailing soils was carried out to investigate metal content and fractionation after leaching. The soil column was divided into four layers after leaching. Fractionation of Cu, Pb, Zn, and Mn in soil was analyzed using a modified BCR sequential extraction method. Metal contents (Cu, Pb, Ca, Mn) in soil layers increased with the depth of the soil column after leaching in the [S, S]-EDDS treatment. The cumulative extraction efficiency was approximately 43.1% for Cu, 26.8% for Zn, 19.5% for Pb, 10.5% for Ca, 2.07% for Mg, 58.5% for Mn, and 7.92% for Fe. The removal of the reducible fractions of Cu and Mn and the exchangeable fraction of Zn was the most significant in the treatment with [S, S]-EDDS. The exchangeable fraction of Pb was the main fraction that was affected by leaching using [S, S]-EDDS. Distribution of Cu and Mn were severely modified by leaching with [S, S]-EDDS. Percentages of residual fractions of the tested heavy metals in the treatment with [S, S]-EDDS after leaching were much higher than that in the control. Although column leaching using [S, S]-EDDS could remove target metals effectively and impaired their availability, it also dissolved large amounts of major elements and modified the distribution of Mn appreciably.  相似文献   

4.
Anaerobic decarboxylation of malonate to acetate was studied withSporomusa malonica, Klebsiella oxytoca, andRhodobacter capsulatus. WhereasS. malonica could grow with malonate as sole substrate (Y=2.0 g·mol–1), malonate decarboxylation byK. oxytoca was coupled with anaerobic growth only in the presence of a cosubstrate, e.g. sucrose or yeast extract (Y s =1.1–1.8 g·mol malonate–1).R. capsulatus used malonate anaerobically only in the light, and growth yields with acetate and malonate were identical. Malonate decarboxylation in cell-free extracts of all three bacteria was stimulated by catalytic amounts of malonyl-CoA, acetyl-CoA, or Coenzyme A plus ATP, indicating that actually malonyl-CoA was the substrate of decarboxylation. Less than 5% of malonyl-CoA decarboxylase activity was found associated with the cytoplasmic membrane. Avidin (except forK. oxytoca) and hydroxylamine inhibited the enzyme completely, EDTA inhibited partially. InS. malonica andK. oxytoca, malonyl-CoA decarboxylase was active only after growth with malonate; malonyl-CoA: acetate CoA transferase was found as well. These results indicate that malonate fermentation by these bacteria proceedsvia malonyl-CoA mediated by a CoA transferase and that subsequent decarboxylation to acetyl-CoA is catalyzed, at least withS. malonica andR. capsulatus, by a biotin enzyme.Abbreviations CoASH Coenzyme A - EDTA ethylenediamine tetraacetate  相似文献   

5.
(2S,3S)-3-methyl- and 3-isopropylaspartic acids were synthesized by bioconversion of the corresponding alkylfumarates (mesaconate and 3-isopropylfumarate) using β-methylaspartase from cell-free extracts of Clostridium tetanomorphum. Optically pure (2S,3S)-3-alkylaspartic acids were transformed in several steps to benzyl (3S,4R)-3-alkylmalolactonates without any racemization of the two chiral centers. These optically active α,β-substituted-β-lactones were polymerized by anionic ring opening polymerization yielding optically active semi-crystalline polyesters. 13C NMR analysis of poly[benzyl β-3-isopropylmalate] in CDCl3 has shown that only the iso-type stereosequence is present in the polymer, indicating that the macromolecular chain is constituted by the only units of benzyl β-(2S,3S)-3-isopropylmalate monomer. The polymerization reaction was done without any racemization of the two stereogenic centers as in the case of benzyl (3S,4R)-3-methylmalolactonate. © 1996 Wiley-Liss, Inc.  相似文献   

6.
7.
Summary A thermo-stable dehalogenase was demonstrated in the crude extracts fromPseudomonas sp (19S). The ability of the enzyme to catalyze the dehalogenation of various halogen-substituted organic acids was investigated and the highest activity was found with monochloroacetate. The enzyme followed Michaelis-Menten kinetics, and the Km for monochloroacetate was 0.2mM. Maximum activity was found at pH 10.5 and 60°C. The enzyme activity in the cell-free extract was unaffected by EDTA or by Mn, Zn, or Cu ions, but was dramatically reduced by HgCl2 (70%) and Pb (NO3)2 (80%).  相似文献   

8.
B. Kos  D. Leštan 《Plant and Soil》2004,263(1):43-51
The feasibility of combined phytoextraction and in situ washing of soil contaminated with Pb (1750 mg kg–1), Zn (1300 mg kg–1), and Cd (7.2 mg kg–1), induced by the addition of biodegradable chelator, [S,S] stereoisomere of ethylenediamine discuccinate ([S,S]-EDDS), was tested in soil columns with hemp (Cannabis sativa) as the phytoextracting plant. The addition of [S,S]-EDDS (10 mmol kg–1 dry soil) yielded concentrations of 1026±442, 330.3±114.7 and 3.84±1.55 mg kg–1 of Pb, Zn and Cd in the dry above-ground plant biomass, respectively. These concentrations were 1926, 7.5, and 11 times higher, respectively, compared to treatments with no chelator addition. Horizontal permeable barriers, composed of a 3 cm high layer of nutrient enriched sawdust and vermiculite mixture, and a 3 cm layer of soil, vermiculite and apatite mixture, were positioned 20, 30 and 40 cm deep in the soil. In chelator treatments, barriers placed 30 cm deep reduced leaching of Pb, Zn and Cd by 435, 4 and 53 times, respectively, compared to columns with no barrier, where 3.0, 4.3 and 3.3% of total initial Pb, Zn and Cd, respectively, was leached during 6-weeks water irrigation after chelator addition. Lower positioned barriers were almost equally effective in preventing leaching of Pb than barriers positioned closer to the soil surface, less effective for Cd, and did not prevent leaching of Zn. 2.53% of total initial Pb and 2.83% of Cd was washed from the contaminated soil and accumulated into the barrier. Combined method was less effective than simulated ex situ soil washing, where 14.2, 5.5 and 24.5% of Pb, Zn and Cd, respectively, were removed after 1-h extraction, but comparable effective to 48-h extraction. Abbreviations: BCF – bioconcentration factor; EDTA – ethylene diaminetetraacetate; HM – heavy metal; PP – phytoextraction potential; [S,S]-EDDS – [S,S]-ethylenediamine disuccinate.  相似文献   

9.
The synthesis of a diastereomerically pure 10-hydroxywarfarin [4-hydroxy-3-(2-hydroxy-3-oxo-1-phenylbutyl)-2H-1 benzopyran-2-one] was accomplished in three steps from racemic warfarin. The relative configuration of the synthetic product was established by conversion to a cyclic derivative followed by NMR and X-ray diffraction analysis. Absolute stereochemistry was determined by enzymatic conversion of either of the pure enantiomers of warfarin to a 10-hydroxy metabolite of known relative configuration. Metabolic formation of 10-hydroxywarfarin was studied using hepatic microsomal preparations from female rats and man. The formation of 10-hydroxywarfarin catalyzed by hepatic microsomes from both dexamethasone-treated rats and man was highly stereoselective [(R)/(S): 3.4-9.0] for (R)-warfarin. In contrast, little stereoselectivity was observed in reactions catalyzed by untreated rat liver microsomes. The resultant stereochemistry at the site of oxidation was also found to be highly dependent on substrate stereochemistry. (R)-Warfarin gave (9R;10S)-10-hydroxywarfarin with only a trace of the (9R;10R) isomer irrespective of which enzyme preparation was used for catalysis, while (S)-warfarin gave (9S;10R)-10-hydroxywarfarin with only a trace of the (9S;10S) isomer, again irrespective of which enzyme preparation was used for catalysis.  相似文献   

10.
The addition of spinach chloroplast total RNA to cell-free extracts from Escherichia coli stimulates amino acid incorporation into protein. The products were characterized by sodium dodecyl sulphate-polyacrylamide gel electrophoresis, and were qualitatively and quantitatively similar to those synthesized in intact isolated chloroplasts. There are two major discrete products of both systems with molecular weights of 52,000 and 35,000. The [35S]methionine-containing chymotryptic peptides of the 52,000 Mr polypeptide synthesized in the E. coli cell-free system have been compared with those of fraction I protein large subunit labelled with [35S]methionine in vivo. From the close similarity in chromatographic properties of the peptides of the two polypeptides, we conclude that the 52,000 Mr product of chloroplast RNA-directed protein synthesis in E. coli extracts is the large subunit of fraction I protein.  相似文献   

11.

Aims

Many studies have proved that EDTA (ethylenediaminetetraacetic acid), EDDS ([S, S’]-ethylenediamine disuccinic acid), and other chelating agents significantly enhance phyto-extraction of copper (Cu) from soil. However, some key factors, such as changes in membrane permeability of root cells and subcellular distribution of Cu and Cu-EDDS complex in leaves and roots, remain unresolved.

Methods

A pot-culture experiment was conducted using soil artificially contaminated with Cu to different degrees to compare its effect on the above factors and the relationship between them in maize (Zea mays L.).

Results

Treatment with 0.5–6.0?mmol?kg?1 (soil) EDDS increased membrane permeability in root cells significantly (p?<?0.05). Chelated Cu accounted for 14.6%–17.4% of the total Cu content of roots and 77.7%–78.8% of that of leaves and was distributed mainly in cell walls in both.

Conclusions

EDDS increases Cu accumulation in shoots mainly by increasing the content of soluble Cu in soil and membrane permeability of root cells. Cu in soil may be absorbed through the apoplastic pathway into the root xylem, translocated to the shoots, and accumulated there as a Cu-EDDS complex.  相似文献   

12.
A 40-day incubation experiment was carried out in order to evaluate the microbial activities and heavy metal availability in long-term contaminated arable and grassland soils after addition of EDTA (ethylenediaminetetraacetic acid) or EDDS ([S,S]-ethylenediaminedisuccinic acid). Soils with similar contamination of heavy metal from the vicinity of a lead smelter were used in the experiment. The soil microbial carbon (Cmic) decreased significantly after addition of EDTA in the arable soil (CM1); lesser effects were observed in the grassland soil (CM2). Addition of EDDS caused a decrease of Cmic during the first 10 days of incubation. In the later phases of the experiment, Cmic increased, and even exceeded the amounts found in the control soils. Respiratory activities and metabolic quotients (qCO2) increased after the addition of the chelating agents into the soils. Higher respiratory activities and qCO2 were observed in the EDTA-treated soils. The readily available heavy metal fractions were extracted with NH4NO3 solution. Readily mobilizable heavy metal fractions of Cd, Pb, Zn, and (in part) Cu increased during the first 3-10 days of incubation in the presence of EDTA. The addition of EDDS particularly increased concentrations of available Cu. Significant correlations between NH4NO3-extractable metals, soil respiratory activities, and qCO2 were found in both soil treatments with EDTA and EDDS. This indicates that enhanced metal mobility seriously affects the microbial processes in experimental soils. In addition, the relationships between NH4NO3-extractable Cd, Cu, and the microbial biomass were found in the CM1 soil amended with EDTA.  相似文献   

13.
Nucleotides such as GTP and GDP appear to be involved in signal transduction via G protein modulation of adenylate cyclase activity. Studies on direct binding of [3H]GDP to membranes prepared from cultured immature rat Sertoli cells indicated that this process was reversible, approached steady state within 10 min, had a Ka of 4.5 ·106M−1 and was specific for guanine nucleotides. The non-hydrolyzable analog, guanosine 5′-O-[3-thio]triphosphate (GPPP[S]), was most effective as an inhibitor of [3H]GDP binding (ED50 = 4.8·10−8M), whereas guanosine 5′-O-[2-thio]diphosphate (Gpp[S]) was less potent (ED50 = 3.4·10−7M). Release of bound GDP was enhanced by follitropin (FSH) in the presence of Gppp[S], although not by FSH alone. Sertoli cell membranes possess guanine nucleotide hydrolase activity, where 95% of added nucleotide was rapidly degraded to guanosine. Binding kinetics were significantly influenced by nucleotide metabolism, which was prevented by controlling the Mg2+ concentration with EDTA and including App[NH]p to reduce nonspecific hydrolysis. Kinetic studies indicated that Gpp[S] inhibited (P < 0.05) Gppp[S]-stimulated adenylate cyclase activity (Ki = 1.8·10−7M), whereas basal activity remained unaffected. Addition of Gpp[S] to pre-activated enzyme (FSH plus GTP) resulted in a time-dependent decay of adenylate cyclase activity with a Koff value of 6 ± 1·min−1. Using a two-stage pre-inculbation technique, adenylate cyclase activity was demonstrated to be sensitive to the nucleotide bound. When FSH was included, catalytic activity was not altered by the order of pre-incubation with the nucleotides. This suggested that the exchange of bound Gpp[S] for Gppp[S] was enhance by FSH. Activation and attenuation of FSH-sensitive adenylate cyclase activity is dependent on a nucleotide exchange mechanism which is driven by (1) the higher affinity of G for GTP than GDP, (2) enhanced release of GD when FSH is present and (3) GTP hydrolysis coupled to rapid metabolism of guanine nucleotides.  相似文献   

14.
The tumorous Drosophila melanogaster blood cell line BII has been studied for evidence for the presence of ecdysteroid receptors. The [3H]ponasterone A (pon A)* used in this study has been extensively purified, and the location of the tritium in the molecule has been partially determined. BII cells do not metabolise ecdysteroids. Intact cells demonstrate a considerable specific uptake of [3H]pon A which is saturable, apparently showing two specific components: a very high affinity component (KD = 0.3 nM) and a high affinity component (KD = 2 nM). The specific binding of [3H]pon A to whole cells is compatible with unlabelled ecdysteroids, but not with mammalian steroid hormones. The association rate constant (ka) for [3H]pon A was determined to be 3 × 107M?1min?1 at 21 °C, while the dissociation rate constant (kd) for the specifically bound [3H]pon A was found to be 4.4 × 10?3/min. Together, the kinetic rate constants yield a value of 0.15 nM for the KD. The receptors have been partially characterised in a cell-free extract prepared by sonification of the cells. The optimum pH for extraction and hormone binding is 8.2. Scatchard plots of binding data indicate that the cell-free extract also contains two high affinity specific binding components (kD = 0.1 nM and KD = 1 nM). The hgih affinity binders are macromolecular, as shown by chromatography on Sephadex G-25, and are susceptible to protease digestion, heat, and treatment with N-ethylmaleimide. Sucrose density centrifugation of the labelled receptor shows one peak at approximately 6S. The stability of the receptor preparation has been studied and conditions have been empirically determined (10% w/v sucrose, 25 mM dithioerthreitol, and 10 mM citrate), whereby the binding capacity of the unlabelled receptor is stable for at least 8 weeks if frozen at ?20°C.  相似文献   

15.
The anthranoid skeleton is believed to be formed by octaketide synthase (OKS), a member of the type III polyketide synthase (PKS) superfamily. Recombinant OKSs catalyze stepwise condensation of eight acetyl units to form a linear octaketide intermediate which, however, is incorrectly folded and cyclized to give the shunt products SEK4 and SEK4b. Here we report in vitro formation of the anthranoid scaffold by cell-free extracts from yeast-extract-treated Cassia bicapsularis cell cultures. Unlike field- and in vitro-grown shoots which accumulate anthraquinones, cell cultures mainly contained tetrahydroanthracenes, formation of which was increased 2.5-fold by the addition of yeast extract. The elicitor-stimulated accumulation of tetrahydroanthracenes was preceded by an approx. 35-fold increase in OKS activity. Incubation of cell-free extracts from yeast-extract-treated cell cultures with acetyl-CoA and [2-14C]malonyl-CoA led to formation of torosachrysone (tetrahydroanthracene) and emodin anthrone, beside two yet unidentified products. No product formation occurred in the absence of acetyl-CoA as starter substrate. To confirm the identities of the enzymatic products, cell-free extracts were incubated with acetyl-CoA and [U-13C3]malonyl-CoA and 13C incorporation was analyzed by ESI-MS/MS. Detection of anthranoid biosynthesis in cell-free extracts indicates in vitro cooperation of OKS with a yet unidentified factor or enzyme for octaketide cyclization.  相似文献   

16.
The influence of metal ions on the metabolism of ethylenediaminetetraacetate (EDTA) by whole cells and cell-free extracts of strain BNC1 was investigated. Metal-EDTA chelates with thermodynamic stability constants below 1012 were readily mineralized by whole cells with maximum specific turnover rates of 15 (MnEDTA) to 20 (Ca-, Mg-, and BaEDTA) μmol g protein−1 min−1. With the exception of ZnEDTA, chelates with stability constants greater than 1012 were not oxidized at a significant rate. However, it was shown for Fe(III)EDTA that even strong complexes can be degraded after pretreatment by addition of calcium and magnesium salts in the pH range 9–11. The range of EDTA chelates converted by cell-free extracts of strain BNC1 did not depend on their thermodynamic stabilities. The EDTA chelates of Ba2+, Co2+, Mg2+, Mn2+, and Zn2+ were oxidized whereas Ca-, Cd-, Cu-, Fe-, Pb-, and SnEDTA were not. The first catabolic enzyme appears to be an EDTA monooxygenase since it requires O2, NADH, and FMN for its activity and yields glyoxylate and ethylenediaminetriacetate as products. The latter is further degraded via N,N′-ethylenediaminediacetate. The maximum specific turnover rate with MgEDTA, the favoured EDTA species, was 50–130 μmol g protein−1 min−1, and the K m value was 120 μmol/l (K s for whole cells = 8 μmol/l). Whole cells as well as cell-free extracts of strain BNC1 also converted several structural analogues of EDTA. Received: 4 July 1997 / Received revision: 25 September 1997 / Accepted: 29 September 1997  相似文献   

17.
Farnesyl pyrophosphate-[14C] and geranylgeranyl pyrophosphate-[14C] were biosynthesized from mevalonic acid-[2-14C] by cell-free enzyme extracts of pea (Pisum sativum) cotyledons containing MgCl2, MnCl2, ATP and AMO-1618. Maximum yields of farnesyl pyrophosphate were obtained after 30 min incubation while geranylgeranyl pyrophosphate was the primary product after 180 min. Biosynthesized geranylgeranyl pyrophosphate-[14C] served as an efficient substrate for ent-kaurene biosynthesis in reaction mixtures containing cotyledon enzymes when AMO-1618 was omitted. Enzyme extracts from green pea shoot tips and chloroplasts also converted geranylgeranyl pyrophosphate to ent-kaurene in very low yields. Ent-kaurene production from mevalonic acid-[2-14C] in extracts of pea shoot tips was also enhanced by addition of chloroplast enzymes. This evidence indicates that kaurene synthetase is present in pea chloroplasts and adds to the possibility that some gibberellin biosynthesis may be compartmentalized in those organelles.  相似文献   

18.
Purple nonsulfur bacteria, Rhodospirillum rubrum and Rhodopseudomonas spheroides were found to possess coenzyme B12-dependent glutamate mutase activity. Cell-free extracts of these bacteria grown on Co2+-containing media catalyzed the conversion of glutamate to β-methylaspartate and further to mesaconate. The activity of the cell-free extracts of these organisms cultivated on Co2+-deficient media was markedly lower than that of the normal cells. Addition of coenzyme B12 to the former reaction mixture enhanced the mesaconate formation via β-methylaspartate. These results indicate the involvement of coenzyme Independent glutamate mutase of these bacteria in the dissimilation of glutamate to acetyl-CoA and pyruvate through the following pathway.

glutamate→β→methylaspartate→mesaconate→citramalate→→acetyl-CoA, pyruvate On the other hand, a greater part of glutamate was converted to α-hydroxyglutarate and succinate with the cell-free extracts of these photosynthetic bacteria. This fact, taking account of the presence of propionyl-CoA carboxylase in these bacteria, implies the participation of coenzyme B12-dependent (R)-methylmalonyl-CoA mutase in the formation of succinate via the following route.

glutamate→α-ketoglutarate→α-hydroxyglutarate→propionate→propionyl-CoA→(S)-methylmalonyl-CoA→(R)-methylmalonyl-CoA→succinyl-CoA  相似文献   

19.
20.
The phytopathogen Rhizoctonia leguminicola has previously been shown to incorporate pipecolic acid into the piperidine alkaloids 1-acetoxy-6-aminooctahydroindolizine (slaframine) and 3,4,5-trihydroxyoctahydro-1-pyrindine. In the experiments described here, resting cultures of R. leguminicola were incubated with [1-14C]- and [2-14C]malonic acid and with [1-14C]- and [2-2H]acetic acid. Both acids were incorporated into the ring systems of both alkaloids. Mass spectrometric analysis of 2H-enriched slaframine showed that the label resides in the five-membered ring and that the methyl carbon of acetate is joined to the carboxyl carbon of pipecolate. A pipecolate-dependent decarboxylation of [1-14C]malonate was demonstrated in cell-free extracts of R. leguminicola. The results account for previously unattributed carbons in the two alkaloids and suggest the formation of an eight-carbon intermediate common to both alkaloids by acylation of malonate with pipecolic acid.  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