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.
A stress-responsive gene, yggG, was introduced into an l-phenylalanine producer, Escherichia coli AJ12741. In shake-flask culture, the yggG-containing recombinant strain (named AJ12741/pHYGG) produced 6.4 g l-phenylalanine l−1 at the end of culture and its yield on glucose was 0.16 g l-phenylalanine g glucose−1. These values are much higher than those of the original AJ12741 strain (3.7 g l-phenylalanine l−1 and 0.09 g l-phenylalanine g glucose−1, respectively). On the other hand, AJ12741/pHYGG strain produced only 4.5 g acetic acid l−1 and its yield on glucose was about a half of that of the AJ12741 culture. Analysis of gene expression revealed that in late growth phase, the expression levels of genes involved in acetic acid production (pta, ackA, and poxB) were relatively low in AJ12741/pHYGG cells. In particular, the level of poxB expression in AJ12741/pHYGG strains was one-seventh of that of the original strain. These results suggest that the formation of a bottleneck for acetic acid production brings about a metabolic flow favorable to l-phenylalanine synthesis in the recombinant strain over-expressing the yggG gene. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

2.
The initial rate ofd-glucosamine uptake by the non-halotolerant yeastSaccharomyces cerevisiae was approximately halved as the apparent half saturation constant (Km) and the apparent maximum velocity (Vmax) changed from 6.6mm to 16.4mm and from 22 μmol · g−1 · min−1 to 16 μmol · g−1 · min−1, respectively, when the salinity in the medium was increased from zerom to 0.68m NaCl. Corresponding changes in a high affinity transport system in the halotolerant yeastDebaryomyces hansenii were from 1.1mm to 4.6mm and from 3.1 μmol · g−1 · min−1 to 4.5 μmol · g−1 · min−1, implying a practically unchanged transport capacity. In 2.7m NaCl, Km and Vmax in this system were 24.5mm and 1.1 μmol · g−1 · min−1, respectively, representing a marked decrease in transport capability. Nevertheless, the degree of affinity in this extreme salinity must still be regarded as noteworthy. In addition to the high affinity transport system inD. hansenii, a low affinity system, presumably without relevance ind-glucosamine transport, was observed.  相似文献   

3.
l-3,4-dihydroxyphenylalanine (l-DOPA) is an aromatic compound employed for the treatment of Parkinson's disease. Metabolic engineering was applied to generate Escherichia coli strains for the production of l-DOPA from glucose by modifying the phosphoenolpyruvate:sugar phosphotransferase system (PTS) and aromatic biosynthetic pathways. Carbon flow was directed to the biosynthesis of l-tyrosine (l-Tyr), an l-DOPA precursor, by transforming strains with compatible plasmids carrying genes encoding a feedback-inhibition resistant version of 3-deoxy-d-arabino-heptulosonate-7-phosphate synthase, transketolase, the chorismate mutase domain from chorismate mutase-prephenate dehydratase from E. coli and cyclohexadienyl dehydrogenase from Zymomonas mobilis. The effects on l-Tyr production of PTS inactivation (PTS gluc+ phenotype), as well as inactivation of the regulatory protein TyrR, were evaluated. PTS inactivation caused a threefold increase in the specific rate of l-Tyr production (q l-Tyr), whereas inactivation of TyrR caused 1.7- and 1.9-fold increases in q l-Tyr in the PTS+ and the PTS gluc+ strains, respectively. An 8.6-fold increase in l-Tyr yield from glucose was observed in the PTS gluc+ tyrR strain. Expression of hpaBC genes encoding the enzyme 4-hydroxyphenylacetate 3-hydroxylase from E. coli W in the strains modified for l-Tyr production caused the synthesis of l-DOPA. One of such strains, having the PTS gluc+ tyrR phenotype, displayed the best production parameters in minimal medium, with a specific rate of l-DOPA production of 13.6 mg/g/h, l-DOPA yield from glucose of 51.7 mg/g and a final l-DOPA titer of 320 mg/l. In a batch fermentor culture in rich medium this strain produced 1.51 g/l of l-DOPA in 50 h.  相似文献   

4.
Bacterial resistance to mercury (Hg) was investigated in strains isolated from Hg-contaminated tailing ponds located in the gold mining area of El Callao (Bolívar State, Venezuela). High frequencies of resistance were detected to both inorganic-Hg and organomercurials among these strains. A broad range of resistance levels was observed when determining minimal inhibitory concentrations of Hg2+. Some strains were able to grow in liquid medium containing 25 μM Hg2+, whereas others grew at 300 μM Hg2+. Of 190 Hg-resistant strains tested, 58.2% were additionally shown to be resistant to ampicillin (40 mg/L), 33.3% to chloramphenicol (30 mg/L), 24.9% to streptomycin (30 mg/L), 23.3% to tetracycline (30 mg/L), and 1.6% to kanamycin (30 mg/L). Furthermore, we found that 20% of the Hg-resistant strains were simultaneously resistant to as many as four of these antibiotics, at the concentrations tested. The presence of large plasmids in 62.9% of 53 Hg-resistant strains screened prompted us to investigate the horizontal transfer of resistance determinants. Mating experiments were performed using Escherichia coli and Pseudomonas aeruginosa as recipient strains. The results obtained confirmed that indigenous Hg-resistant bacteria colonizing the tailing ponds can effectively transfer the phenotype to potentially pathogenic species.  相似文献   

5.
We evaluated phytohormone and polyamine biosynthesis, siderophore production, and phosphate solubilization in two strains (Cd and Az39) of Azospirillum brasilense used for inoculant formulation in Argentina during the last 20 years. Siderophore production and phosphate solubilization were evaluated in a chemically defined medium, with negative results. Indole 3-acetic acid (IAA), gibberellic acid (GA3), and abscisic acid (ABA) production were analyzed by gas chromatography-mass spectrometry. Ethylene, polyamine, and zeatin (Z) biosynthesis were determined by gas chromatography-flame ionization detector and high performance liquid chromatography (HPLC-fluorescence and -UV), respectively. Phytohormones IAA, Z, GA3, ABA, ethylene, and growth regulators putrescine, spermine, spermidine, and cadaverine (CAD) were found in culture supernatant of both strains. IAA, Z, and GA3 were found in all two strains; however, their levels were significantly higher (p < 0.01) in Cd (10.8, 2.32, 0.66 μg ml−1). ABA biosynthesis was significantly higher (p < 0.01) in Az39 (0.077 μg ml−1). Ethylene and polyamine CAD were found in all two strains, with highest production in Cd cultured in NFb plus l-methionine (3.94 ng ml−1 h−1) and Az39 cultured in NFb plus l-lysine (36.55 ng ml−1 h−1). This is the first report on the evaluation of important bioactive molecules in strains of A. brasilense as potentially capable of direct plant growth promotion or agronomic yield increase. Az39 and Cd showed differential capability to produce the five major phytohormones and CAD in chemically defined medium. This fact has important technological implications for inoculant formulation as different concentrations of growth regulators are produced by different strains or culture conditions.  相似文献   

6.
When ruminal fluid from a cow fed timothy hay was serially diluted (10-fold increments into anaerobic broth containing 15 mg ml−1 Trypticase), the low dilutions (≤10−6) had optical densities greater than 2.0 and ammonia concentrations greater than 100 mM. The optical densities and ammonia concentrations of the 10−8 and 10−9 dilutions were very low, but large cocci were observed in the 10−8 dilution. The large cocci were isolated and identified by 16S rDNA sequencing as Megasphaera elsdenii. The freshly isolated strain (JL1) grew well on Trypticase, but less than 4% of the amino acid nitrogen in Trypticase was converted to ammonia. Optical density and ammonia production were twice as great if Casamino acids were provided, and similar results were obtained with seven other strains (B159, AW106, YT91, LC1, T81, J1, and YZ70). Specific activities of deamination (based on Casamino acids) of the eight strains ranged from 100 (strain JL1) to 325 (strain B159) nmol mg protein−1 min−1. None of the strains could utilize branched-chain amino acids as an energy source for growth, but specific activities of branched-chain amino acid deamination ranged from 15 to 65 nmol mg protein−1 min−1. All eight of the M. elsdenii strains grew well in the presence of 5 μM monensin, and only two of the strains were strongly inhibited by 20 μM monensin. On the basis of these results, it appears that M. elsdenii is deficient in peptidase activity and can utilize only a few amino acids. Some M. elsdenii strains produced ammonia and branched-chain volatile fatty acids nearly as fast as obligate amino acid-fermenting ruminal bacteria, but the extent of this production was at least fourfold lower. Because all of the strains could tolerate 5 μM monensin, it is unlikely that this feed additive would significantly inhibit M. elsdenii in vivo. Received: 12 December 2001 / Accepted: 5 February 2002  相似文献   

7.
Escherichia coli K12 strains producing l-phenylalanine were converted to l-tyrosine-producing strains using a novel genetic method for gene replacement. We deleted a region of the E. coli K12 chromosome including the pheA gene encoding chorismate mutase/prephenate dehydratase, its leader peptide (pheL), and its promoter using a new polymerase chain reaction-based method that does not leave a chromosomal scar. For high level expression of tyrA, encoding chorismate mutase/prephenate dehydrogenase, its native promoter was replaced with the strong trc promoter. The linked ΔpheLA and Ptrc-tyrA::KanR genetic modifications were moved into l-phenylalanine producing strains by generalized transduction to convert l-phenylalanine-producing strains to l-tyrosine-producing strains. Moreover, introduction of a plasmid carrying genes responsible for sucrose degradation into these strains enabled l-tyrosine-production from sucrose.  相似文献   

8.
l-Ribose is a rare and expensive sugar that can be used as a precursor for the production of l-nucleoside analogues, which are used as antiviral drugs. In this work, we describe a novel way of producing l-ribose from the readily available raw material l-arabinose. This was achieved by introducing l-ribose isomerase activity into l-ribulokinase-deficient Escherichia coli UP1110 and Lactobacillus plantarum BPT197 strains. The process for l-ribose production by resting cells was investigated. The initial l-ribose production rates at 39°C and pH 8 were 0.46 ± 0.01 g g−1 h−1 (1.84 ± 0.03 g l−1 h−1) and 0.27 ± 0.01 g g−1 h−1 (1.91 ± 0.1 g l−1 h−1) for E. coli and for L. plantarum, respectively. Conversions were around 20% at their highest in the experiments. Also partially purified protein precipitates having both l-arabinose isomerase and l-ribose isomerase activity were successfully used for converting l-arabinose to l-ribose.  相似文献   

9.
Chloride (Cl) conductances were studied in primary cultures of the bright part of rabbit distal convoluted tubule (DCTb) by the whole cell patch clamp technique. The bath solution (33°C) contained (in mm): 140 NaCl, 1 CaCl2, 10 N-2-hydroxy-ethylpiperazine-N′-2-ethanesulfonic acid (HEPES), pH 7.4 and the pipette solution 140 N-methyl-d-glucamine (NMDG)-Cl, 5 MgATP, 1 ethylene-glycol-bis(b-aminoethyl ether)-N,N,N,N′-tetraacetic acid (EGTA), 10 HEPES, pH 7.4. We identified a Cl current activated by 10−5 m forskolin, 10−3 m 8-bromo adenosine 3′,5′-cyclic monophophosphate (8 Br-cAMP), 10−6 m phorbol 12-myristate 13-acetate (PMA), 10−3 m intracellular adenosine 3′,5′-cyclic monophophosphate (cAMP) and 10−7 m calcitonin. The current-voltage relationship was linear and the relative ion selectivity was Br > Cl≫ I > glutamate. This current was inhibited by 10−3 m diphenylamine-2-carboxylate (DPC) and 10−4 m 5-nitro-2-(3-phenylpropylamino)-benzoate (NPPB) and was insensitive to 10−3 m 4,4′-diisothiocyanostilbene-2,2′-disulfonic acid (DIDS). These characteristics are similar to those described for the cystic fibrosis transmembrane conductance regulator (CFTR) Cl conductance. In a few cases, forskolin and calcitonin induced an outwardly rectifying Cl current blocked by DIDS. To determine the exact location of the Cl conductance 6-methoxy-1-(3-sulfonatopropyl) quinolinium (SPQ) fluorescence experiments were carried out. Cultures seeded on collagen-coated permeable filters were loaded overnight with 5 mm SPQ and the emitted fluorescence analyzed by laser-scan cytometry. Cl removal from the apical solution induced a Cl efflux which was stimulated by 10−5 m forskolin, 10−7 calcitonin and inhibited by 10−5 m NPPB. In 140 mm NaBr, forskolin stimulated an apical Br influx through the Cl pathway. Forskolin and calcitonin had no effect on the basolateral Cl permeability. Thus in DCTb cultured cells, exposure to calcitonin activates a Cl conductance in the apical membrane through a cAMP-dependent mechanism. Received: 5 July 1995/Revised: 21 December 1995  相似文献   

10.
The aim of this work was to evaluate phytohormone biosynthesis, siderophores production, and phosphate solubilization in three strains (E109, USDA110, and SEMIA5080) of Bradyrhizobium japonicum, most commonly used for inoculation of soybean and nonlegumes in USA, Canada, and South America. Siderophore production and phosphate solubilization were evaluated in selective culture conditions, which had negative results. Indole-3-acetic acid (IAA), gibberellic acid (GA3), and abscisic acid (ABA) production were analyzed by gas chromatography–mass spectrometry (GC-MS). Ethylene and zeatin biosynthesis were determined by GS–flame ionization detection and high-performance liquid chromatography (HPLC-UV), respectively. IAA, zeatin, and GA3 were found in all three strains; however, their levels were significantly higher (p < 0.01) in SEMIA5080 (3.8 μg ml−1), USDA110 (2.5 μg ml−1), and E109 (0.87 μg ml−1), respectively. ABA biosynthesis was detected only in USDA110 (0.019 μg ml−1). Ethylene was found in all three strains, with highest production rate (18.1 ng ml−1 h−1) in E109 cultured in yeast extract mannitol medium plus l-methionine. This is the first report of IAA, GA3, zeatin, ethylene, and ABA production by B. japonicum in pure cultures, using quantitative physicochemical methodology. The three strains have differential capability to produce the five major phytohormones and this fact may have an important technological implication for inoculant formulation.  相似文献   

11.
Biosynthesis of six saponins (ginsenosides) in suspension culture of P. quinquefolium Z5 was investigated. Ginsenoside content in biomass reached the highest level, nearly 30 mg g−1 d.w., between 25 and 30 days of the culture. Saponins were synthesized simultaneously with cell growth but their synthesis rate was not proportional to the growth rate. During the phase of rapid biomass multiplication, after which biomass reached 90% of its maximum yield, only half examined ginsenosides was produced. The second half of the final saponins yield was produced during the slow growth phase, in which only 10% of biomass was grown. During the intensive growth phase the productivity of six saponins examined per biomass (dry weight) unit was 3.4 μg mg−1 d.w. day−1, however, this parameter calculated for slow growth phase reached nearly 30 μg mg−1 d.w. day−1. There were differences in increase of the contents of six saponins determined in biomass, and it was the highest for saponins Re (20(S)-protopanaxatriol-6-[O-α-l-rhamnopyranosyl(1 → 2)-β-d-glucopyranoside]-20-O-β-d-glucopyranoside) and Rg1 (20(S)-protopanaxatriol-6,20-di-O-β-d-glucoside).  相似文献   

12.
In a study screening anaerobic microbes utilizing d-galactitol as a fermentable carbon source, four bacterial strains were isolated from an enrichment culture producing H2, ethanol, butanol, acetic acid, butyric acid, and hexanoic acid. Among these isolates, strain BS-1 produced hexanoic acid as a major metabolic product of anaerobic fermentation with d-galactitol. Strain BS-1 belonged to the genus Clostridium based on phylogenetic analysis using 16S rRNA gene sequences, and the most closely related strain was Clostridium sporosphaeroides DSM 1294T, with 94.4% 16S rRNA gene similarity. In batch cultures, Clostridium sp. BS-1 produced 550 ± 31 mL L−1 of H2, 0.36 ± 0.01 g L−1 of acetic acid, 0.44 ± 0.01 g L−1 of butyric acid, and 0.98 ± 0.03 g L−1 of hexanoic acid in a 4-day cultivation. The production of hexanoic acid increased to 1.22 and 1.73 g L−1 with the addition of 1.5 g L−1 of sodium acetate and 100 mM 2-(N-morpholino)ethanesulfonic acid (MES), respectively. Especially when 1.5 g L−1 of sodium acetate and 100 mM MES were added simultaneously, the production of hexanoic acid increased up to 2.99 g L−1. Without adding sodium acetate, 2.75 g L−1 of hexanoic acid production from d-galactitol was achieved using a coculture of Clostridium sp. BS-1 and one of the isolates, Clostridium sp. BS-7, in the presence of 100 mM MES. In addition, volatile fatty acid (VFA) production by Clostridium sp. BS-1 from d-galactitol and d-glucose was enhanced when a more reduced culture redox potential (CRP) was applied via addition of Na2S·9H2O.  相似文献   

13.
Xylanase production by seven fungal strains was investigated using concentrated spent sulphite liquor (SSLc), xylan and d-xylose as carbon substrates. An SSLc-based medium induced xylanase production at varying levels in all of these strains, with Aspergillus oryzae NRRL 3485 and Aspergillus phoenicis ATCC 13157 yielding activities of 164 and 146 U ml−1, respectively; these values were higher than those obtained on xylan or d-xylose with the same fungal strains. The highest xylanase activity of 322 U ml−1 was obtained with Aspergillus foetidus ATCC 14916 on xylan. Electrophoretic and zymogram analysis indicated three xylanases from A. oryzae with molecular weights of approximately 32, 22 and 19 kDa, whereas A. phoenicis produced two xylanases with molecular weights of about 25 and 21 kDa. Crude xylanase preparations from these A. oryzae and A. phoenicis strains exhibited optimal activities at pH 6.5 and 5.0 and at 65 and 55°C, respectively. The A. oryzae xylanolytic activity was stable at 50°C over the pH range 4.5–10. The crude xylanase preparations from these A. oryzae and A. phoenicis strains had negligible cellulase activity, and their application in the biobleaching of hardwood pulp reduced chlorine dioxide consumption by 20–30% without sacrificing brightness.  相似文献   

14.
EstA was purified from the supernatant by A. lwoffii 16C-1. Its molecular mass was determined to be 45 kDa, and the optimal activity occurred when the pH level was 8.0 at a temperature of 37°C. The activation energies for the hydrolysis of p-nitrophenyl butyrate was determined to be 11.25 kcal/mol in the temperature range of 10–37°C. The enzyme was unstable at temperatures higher than 50°C. The Michaelis constant (K m ) and V max for p-nitrophenyl butyrate were 11 μM and 131.6 μM min−1 mg of protein-1, respectively. The enzyme was strongly inhibited by Hg2−, Ca2+, Mg2+, Fe2+, Cu2+, Zn2+, Mn2+, Co2+, ethylemediaminetetraacetic acid (EDTA), phenylmethylsulfonyl fluoride (PMSF), and diisopropyl fluorophosphate (DFP). Received: 20 August 2001 / Accepted: 20 September 2001  相似文献   

15.
A novel moderately halophilic, alkaliphilic, non-motile, non-sporulating, catalase-positive, oxidase-negative, aerobic, coccus-shaped, Gram-positive bacterium, designated strain JSM 071043T, was isolated from a subterranean brine sample collected from a salt mine in Hunan Province, China. Growth occurred with 0.5–20% (w/v) NaCl (optimum 5–10%) at pH 6.5–10.5 (optimum pH 8.5) and at 10–40°C (optimum 25–30°C). Good growth also occurred in the presence of 0.5–20% (w/v) KCl (optimum 5–8%) or 0.5–25% (w/v) MgCl2·6H2O (optimum 5–10%). The peptidoglycan type was A4α (l-Lys–l-Ala–l-Glu) and major cell-wall sugars were tyvelose and mannose. The major cellular fatty acids were anteiso-C15:0, iso-C16:0 and anteiso-C17:0. Strain JSM 071043T contained MK-9 and MK-8 as the predominant menaquinones and diphosphatidylglycerol, phosphatidylglycerol and phosphatidylinositol as the major polar lipids. The DNA G + C content was 67.8 mol%. Phylogenetic analysis based on 16S rRNA gene sequences showed that strain JSM 071043T was a member of the suborder Micrococcineae, and was most closely related to Zhihengliuella halotolerans YIM 70185T (sequence similarity 98.9%) and Zhihengliuella alba YIM 90734T (98.2%), and the three strains formed a distinct branch in the phylogenetic tree. The combination of phylogenetic analysis, DNA–DNA relatedness values, phenotypic characteristics and chemotaxonomic data supports the proposal that strain JSM 071043T represents a novel species of the genus Zhihengliuella, for which the name Z. salsuginis sp. nov. is proposed. The type strain is JSM 071043T (= DSM 21149T = KCTC 19466T).  相似文献   

16.
Brevibacterium flavum ATCC14067 was engineered for l-valine production by overexpression of different ilv genes; the ilvEBNrC genes from B. flavum NV128 provided the best candidate for l-valine production. In traditional fermentation, l-valine production reached 30.08 ± 0.92 g/L at 31°C in 72 h with a low conversion efficiency of 0.129 g/g. To further improve the l-valine production and conversion efficiency based on the optimum temperatures of l-valine biosynthesis enzymes (above 35°C) and the thermotolerance of B. flavum, the fermentation temperature was increased to 34, 37, and 40°C. As a result, higher metabolic rate and l-valine biosynthesis enzymes activity were obtained at high temperature, and the maximum l-valine production, conversion efficiency, and specific l-valine production rate reached 38.08 ± 1.32 g/L, 0.241 g/g, and 0.133 g g−1 h−1, respectively, at 37°C in 48 h fermentation. The strategy for enhancing l-valine production by overexpression of key enzymes in thermotolerant strains may provide an alternative approach to enhance branched-chain amino acids production with other strains.  相似文献   

17.
Lipomyces starkeyi is an oleaginous yeast, and has been classified in four distinct groups, i.e., sensu stricto and custers α, β, and γ. Recently, L. starkeyi clusters α, β, and γ were recognized independent species, Lipomyces mesembrius, Lipomyces doorenjongii, and Lipomyces kockii, respectively. In this study, we investigated phylogenetic relationships within L. starkeyi, including 18 Japanese wild strains, and its related species, based on internal transcribed spacer sequences and evaluated biochemical characters which reflected the phylogenetic tree. Phylogenetic analysis showed that most of Japanese wild strains formed one clade and this clade is more closely related to L. starkeyi s.s. clade including one Japanese wild strain than other clades. Only three Japanese wild strains were genetically distinct from L. starkeyi. Lipomyces mesembrius and L. doorenjongii shared one clade, while L. kockii was genetically distinct from the other three species. Strains in L. starkeyi s.s. clade converted six sugars, d-glucose, d-xylose, l-arabinose, d-galactose, d-mannose, and d-cellobiose to produce high total lipid yields. The Japanese wild strains in subclades B, C, and D converted d-glucose, d-galactose, and d-mannose to produce high total lipid yields. Lipomyces mesembrius was divided into two subclades. Lipomyces mesembrius CBS 7737 converted d-xylose, l-arabinose, d-galactose, and d-cellobiose, while the other L. mesembrius strains did not. Lipomyces doorenjongii converted all the sugars except d-cellobiose. In comparison to L. starkeyi, L. mesembrius, and L. doorenjongii, L. kockii produced higher total lipid yields from d-glucose, d-galactose, and d-mannose. The type of sugar converted depended on the subclade classification elucidated in this study.  相似文献   

18.
A ferric reductase was purified into an electrophoretically homologous state from Magnetospirillum gryphiswaldense MSR-1 strain. The enzyme was found within the cytoplasm and associated with the cytoplasmic membrane. The molecular weight of the purified enzyme was calculated as 16.1 kDa using sodium dodecyl sulfate–polyacrylamide gel electrophoresis, and was almost identical to that calibrated using agarose gel filtration chromatography. It was NADH-dependent and required flavin mononucleotide as a cofactor. The optimal reaction temperature and pH values were 30°C and 6.5, respectively. The K m and Vmax values for ferric citrate were 45.1 μM and 1.216 μM min−1, respectively. Though ferric reductase activity could be inhibited by Co2+, Cu2+, Mn2+, and Zn2+, even high concentrations of Mg2+ ions have failed to accomplish such enzyme inhibition. Furthermore, the molecular weight, the N-terminal sequence, and the activity of ferric reductase from MSR-1 are not matching with the enzyme preparation obtained from an analogous strain M. magnetotacticum (MS-1). Therefore, it is concluded that the ferric reductase of M. grysphiwaldense and M. magnetotacticum strains are two different enzymes.  相似文献   

19.
A set of three oxaliplatin derivatives containing 1,2-trans-R,R-diaminocyclohexane (dach) as a spectator ligand and different chelating leaving groups X–Y, viz., [Pt(dach)(O,O-cyclobutane-1,1-dicarboxylate)], or Pt(dach)(CBDCA), [Pt(dach)(N,O-glycine)]+, or Pt(dach)(gly), and [Pt(dach)(N,S-methionine)]+, or Pt(dach)(l-Met), where l-Met is l-methionine, were synthesized and the crystal structure of Pt(dach)(gly) was determined by X-ray diffraction. The effect of the leaving group on the reactivity of the resulting Pt(II) complexes was studied for the nucleophiles thiourea, glutathione (GSH) and l-Met under pseudo-first-order conditions as a function of nucleophile concentration and temperature, using UV–vis spectrophotometric techniques. 1H NMR spectroscopy was used to follow the substitution of the leaving group by guanosine 5′-monophosphate (5′-GMP2−) under second-order conditions. The rate constants indicate for all reactions a direct substitution of the X–Y chelate by the selected nucleophiles, thereby showing that the nature of the chelate, viz., O–O (CBDCA2−), N–O (glycine) or S–N (l-Met), respectively, plays an important role in the kinetic and mechanistic behavior of the Pt(II) complex. The k 1 values for the reaction with thiourea, l-Met, GSH and 5′-GMP2− were found to be as follows (103 k 1, 37.5 °C, M−1 s−1): Pt(dach)(CBDCA) 61 ± 2, 21.6 ± 0.1, 23 ± 1, 0.352 ± 0.002; Pt(dach)(gly) 82 ± 3, 6.2 ± 0.2, 37 ± 1, 1.77 ± 0.01; Pt(dach)(l-Met) (thiourea, GSH) 62 ± 2, 24 ± 1. The activation parameters for all reactions studied suggest an associative substitution mechanism. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

20.
Lens Major Intrinsic Protein (MIP) is a member of a family of membrane transport proteins including the Aquaporins and bacterial glycerol transporters. When expressed in Xenopus oocytes, MIP increased both glycerol permeability and the activity of glycerol kinase. Glycerol permeability (p Gly ) was 2.3 ± 0.23 × 10−6 cm sec−1 with MIP vs. 0.92 ± 0.086 × 10−6 cm sec−1 in control oocytes. The p Gly of MIP was independent of concentration from 5 × 10−5 to 5 × 10−2 m, had a low temperature dependence, and was inhibited approximately 90%, 80% and 50% by 1.0 mm Hg++, 0.2 mm DIDS (diisothiocyanodisulfonic stilbene), and 0.1 mm Cu++, respectively. MIP-enhanced glycerol phosphorylation, resulting in increased incorporation of glycerol into lipids. This could arise from an increase in the total activity of glycerol kinase, or from an increase in its affinity for glycerol. Based on methods we present to distinguish these mechanisms, MIP increased the maximum rate of phosphorylation by glycerol kinase (0.12 ± 0.03 vs. 0.06 ± 0.01 pmol min−1 cell−1) without changing the binding of glycerol to the kinase (K M ∼ 10 μm). Received: 23 May 1997/Revised: 4 August 1997  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