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 15 毫秒
1.
Dmitriy N. Shevela 《BBA》2006,1757(4):253-261
It is shown that the hydrazine-induced transition of the water-oxidizing complex (WOC) to super-reduced S-states depends on the presence of bicarbonate in the medium so that after a 20 min treatment of isolated spinach thylakoids with 3 mM NH2NH2 at 20 °C in the CO2/HCO3-depleted buffer the S-state populations are: 42% of S−3, 42% of S−2, 16% of S−1 and even formal S−4 state is reached, while in the presence of 2 mM NaHCO3, the same treatment produces 30% of S−3, 38% of S−2, and 32% of S−1 and there is no indication of the S−4 state. Bicarbonate requirement for the oxygen-evolving activity, very low in untreated thylakoids, considerably increases upon the transition of the WOC to the super-reduced S-states, and the requirement becomes low again when the WOC returns back to the normal S-states using pre-illumination. The results are discussed as a possible indication of ligation of bicarbonate to manganese ions within the WOC.  相似文献   

2.
It has been shown that thermoinactivation of the isolated D1/D2/cytochrome b 559 complex (RC) of photosystem 2 (PS-2) from pea under anaerobic conditions at 35°C in 20 mM Tris-HCl buffer (pH 7.2) depleted of HCO3, with 35 mM NaCl and 0.05% n-dodecyl-β-maltoside, results in a decrease in photochemical activity measured by photoreduction of the PS-2 primary electron acceptor, pheophytin (by 50% after 3 min of heating), which is accompanied by aggregation of the D1 and D2 proteins. Bicarbonate, formate, and acetate anions added to the sample under these conditions differently influence the maintenance of photochemical activity: a 50% loss of photochemical activity occurs in 11.5 min of heating in the presence of bicarbonate and in 4 and 4.6 min in the presence of formate and acetate, respectively. The addition of bicarbonate completely prevents aggregation of the D1 and D2 proteins as opposed to formate and acetate (their presence has no effect on the aggregation during thermoinactivation). Since the isolated RCs have neither inorganic Mn/Ca-containing core of the water-oxidizing complex nor nonheme Fe2+, it is supposed that bicarbonate specifically interacts with the hydrophilic domains of the D1 and D2 proteins, which prevents their structural modification that is a signal for aggregation of these proteins and the loss of photochemical activity.  相似文献   

3.
In the present study, N and S assimilation, antioxidant enzymes activity, and yield were studied in N and S-treated plants of Brassica juncea (L.) Czern. & Coss. (cvs. Chuutki and Radha) under salt stress. The treatments were given as follows: (1) NaCl90 mM+N0S0 mg kg-1 sand (control), (2) NaCl90 mM+N60S0 mg kg-1 sand, (3) NaCl90 mM+N60S20 mg kg-1 sand, (4) NaCl90 mM+N60S40 mg kg-1 sand, and (5) NaCl90 mM+N60S60 mg kg-1 sand. The combined application of N (60 mg kg−1 sand) and S (40 mg kg−1 sand) proved beneficial in alleviating the adverse effect of salt stress on growth attributes (shoot length plant−1, fresh weight plant−1, dry weight plant−1, and area leaf−1), physio-biochemical parameters (carbonic anhydrase activity, total chlorophyll, adenosine triphosphate-sulphurylase activity, leaf N, K and Na content, K/Na ratio, activity of nitrate reductase, nitrite reductase, glutamine synthetase, glutamate synthase, catalase, superoxide dismutase, ascorbate peroxidase and glutathione reductase, and content of glutathione and ascorbate), and yield attributes (pods plant−1, seeds pod−1, and seed yield plant−1). Therefore, it is concluded that combined application of N and S induced the physiological and biochemical mechanisms of Brassica. The stimulation of antioxidant enzymes activity and its synergy with N and S assimilation may be one of the important mechanisms that help the plants to tolerate the salinity stress and resulted in an improved yield.  相似文献   

4.
Currently available data about bicarbonate (BC) action on the Mn-containing water-oxidizing complex (WOC) of the photosystem II (PSII) were obtained almost solely in vitro, e.g. on subchloroplast membrane fragments enriched with PSII. To investigate the in vivo BC effect on the PSII donor side, we used the method of dark thermoinactivation of intact Chlamydomonas reinhardtii cells. Photosynthetic activity of PSII was measured as photoinduced changes in the PSII chlorophyll fluorescence yield and as the rate of photosynthetic oxygen evolution. To exclude a “direct” effect of the absence of BC on the PSII activity, before measurements of the photosynthetic activity, the concentration of BC in all samples was equalized by addition of NaHCO3 to each of them (except for those that contained 5 mM of NaHCO3 during thermoinactivation) to reach the final concentration of 5 mM. This allowed registering only so-called “irreversible” (i.e., not reversible by subsequent addition of BC) effect of the absence of BC during thermoinactivation. It was shown that, if 5 mM NaHCO3 was added to the medium before thermoinactivation, the rate of inactivation of the PSII donor side was lower than in BC-depleted medium 1.5-to 2-fold. The obtained results are interpreted as an indication that BC protects the donor side of PSII against thermoinactivation in vivo, in intact C. reinhardtii cells. This proves the correctness of the earlier proposition that BC is an integral constituent of the Mn-containing water-oxidizing complex of PSII. Published in Russian in Fiziologiya Rastenii, 2007, Vol. 54, No. 3, pp. 342–349. The article was translated by the authors.  相似文献   

5.
Yuncai Hu  Urs Schmidhalter 《Planta》1998,204(2):212-219
Wheat leaf growth is known to be spatially affected by salinity. The altered spatial distribution of leaf growth under saline conditions may be associated with spatial changes in tissue mineral elements. The objective of this study was to evaluate the spatial distributions of mineral elements and their net deposition rates in the elongating and mature zones of leaf 4 of the main stem of spring wheat (Triticum aestivum L. cv. Lona) during its linear growth phase under saline soil conditions. Plants were grown in an illitic-chloritic silty loam with 0 and 120 mM NaCl. Three days after emergence of leaf 4, sampling was begun at 3 and 13 h into the 16-h light period. Spatial distributions of fresh weight (FW), dry weight (DW), and Na+, K+, Cl, NO 3, Ca2+, Mg2+, total P, and total N in the elongating and mature tissues were determined on a millimeter scale. The patterns of spatial distribution of Na+, Cl, K+, NO3 , and Ca2+ in the growing leaves were affected by salinity, while those of Mg2+, total P, and total N were not. Sodium, K+, Cl, Ca2+, Mg2+, and total N concentrations (mmol · kg−1 FW) were consistently higher at 120 mM NaCl than at 0 mM NaCl along the leaf axis from the leaf base, whereas NO3 concentration was lower at 120 mM NaCl. Deposition rates of all nutrients were greatest in the elongation zone. The elongation zone was the strongest sink for mineral elements in the leaf tissues. Local net deposition rates of Na+, Cl, Ca2+, and Mg2+ (mmol · kg−1 FW · h−1) in the most actively elongating zone were enhanced by 120 mM NaCl, whereas for NO3 this was depressed. The lower supply of NO 3 to growing leaves may be responsible for the inhibition of growth under saline conditions. Higher tissue concentrations of Na+ and Cl may cause ion imbalance but probably did not result in ion toxicity in the growing leaves. Potassium, Ca2+, Mg2+, total P, and total N are less plausibly responsible for the reduction in leaf growth in this study. Higher tissue K+ and Ca2+ concentrations at 120 mM NaCl are probably due to the presence of high Ca2+ in the soil of this study. Received: 13 March 1997 / Accepted: 9 June 1997  相似文献   

6.
The aim of this work was to investigate the possibility of conducting a continuous aerobic bioprocess in a horizontal rotating tubular bioreactor (HRTB). Aerobic oxidation of acetate by the action of a mixed microbial culture was chosen as a model process. The microbial culture was not only grown in a suspension but also in the form of a biofilm on the interior surface of HRTB. Efficiency of the bioprocess was monitored by determination of the acetate concentration and chemical oxygen demand (COD). While acetate inlet concentration and feeding rate influenced efficiency of acetate oxidation, the bioreactor rotation speed did not influence the bioprocess dynamics significantly. Gradients of acetate concentration and pH along HRTB were more pronounced at lower feeding rates. Volumetric load of acetate was proved to be the most significant parameter. High volumetric loads (above 2 g acetate l−1 h−1) gave poor acetate oxidation efficiency (only 17 to 50%). When the volumetric load was in the range of 0.60–1.75 g acetate l−1 h−1, acetate oxidation efficiency was 50–75%. At lower volumetric loads (0.14–0.58 g acetate l−1 h−1), complete acetate consumption was achieved. On the basis of the obtained results, it can be concluded that HRTB is suitable for conducting aerobic continuous bioprocesses.  相似文献   

7.
High-rate biological conversion of sulfide and nitrate in synthetic wastewater to, respectively, elemental sulfur (S0) and nitrogen-containing gas (such as N2) was achieved in an expanded granular sludge bed (EGSB) reactor. A novel strategy was adopted to first cultivate mature granules using anaerobic sludge as seed sludge in sulfate-laden medium. The cultivated granules were then incubated in sulfide-laden medium to acclimate autotrophic denitrifiers. The incubated granules converted sulfide, nitrate, and acetate simultaneously in the same EGSB reactor to S0, N-containing gases and CO2 at loading rates of 3.0 kg S m−3 d−1, 1.45 kg N m−3 d−1, and 2.77 kg Ac m−1 d−1, respectively, and was not inhibited by sulfide concentrations up to 800 mg l−1. Effects of the C/N ratio on granule performance were identified. The granules cultivated in the sulfide-laden medium have Pseudomonas spp. and Azoarcus sp. presenting the heterotrophs and autotrophs that co-work in the high-rate EGSB-SDD (simultaneous desulfurization and denitrification) reactor.  相似文献   

8.
Mechanistic and structural aspects of photosynthetic water oxidation   总被引:10,自引:0,他引:10  
Conclusions on the functional and structural organisation of photosynthetic water oxidation are gathered from a critical survey of the wealth of data reported in the literature and author's own experimental research: (1) the water oxidising complex (WOC) contains a tetranuclear manganese cluster of dimer of dimers' structure and functional heterogeneity of the metal centers, (2) the four step univalent oxidative pathway leading to water oxidation into molecular oxygen and four protons comprises only manganese, tyrosine YZ of polypeptide Dl and the substrate as redox active species, (3) the redox transitions S0→ S1 and S1→ S2 are manganese centered whereas S2→ S3 is most likely a ligand-centered reaction, (4) there exist several lines of evidence for a marked structural change that accompanies the redox transition S2→ S3, (5) one Ca2+ is an indispensible constituent of a functionally competent WOC while the role of Cl is much less clear and a direct participation disputable, (6) substrate water is most likely bound in all redox states S0,…,S3 and exchangeable with the bulk phase. The protonation state is determined by the redox state S1 and the protein microenvironment. A mechanism is proposed for water oxidation in the WOC that is based on three key postulates: (1) water oxidation takes place in the first coordination sphere of one manganese dimer [MnaMnb]; (2) the essential O-O bond is preformed in S3 as part of a rapid redox isomerism S3(I)→S3(II) where in S3(II) a nuclear geometry and electronic configuration is attained that corresponds to a peroxidic-type species; and (3) S3(II) is an ‘entatic state’ for the formation of complexed dioxygen triggered by YZOX induced electron abstraction from the WOC and electronic redistribution to S0(O2).  相似文献   

9.
The biological reduction of Fe(III) ethylenediaminetetraacetic acid (EDTA) is a key step for NO removal in a chemical absorption–biological reduction integrated process. Since typical flue gas contain oxygen, NO2 and NO3 would be present in the absorption solution after NO absorption. In this paper, the interaction of NO2 , NO3 , and Fe(III)EDTA reduction was investigated. The experimental results indicate that the Fe(III)EDTA reduction rate decrease with the increase of NO2 or NO3 addition. In the presence of 10 mM NO2 or NO3 , the average reduction rate of Fe(III)EDTA during the first 6-h reaction was 0.076 and 0.17 mM h−1, respectively, compared with 1.07 mM h−1 in the absence of NO2 and NO3 . Fe(III)EDTA and either NO2 or NO3 reduction occurred simultaneously. Interestingly, the reduction rate of NO2 or NO3 was enhanced in presence of Fe(III)EDTA. The inhibition patterns observed during the effect of NO2 and NO3 on the Fe(III)EDTA reduction experiments suggest that Escherichia coli can utilize NO2 , NO3 , and Fe(III)EDTA as terminal electron acceptors.  相似文献   

10.
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   

11.
Lee S  Kim J  Shin SG  Hwang S 《Biotechnology letters》2008,30(6):1011-1016
The biokinetics of glucose metabolism were evaluated in Aeromonas hydrophila during growth in an anaerobic biosystem. After approx 34 h growth, A. hydrophila metabolized 5,000 mg glucose l−1 into the end-products ethanol, acetate, succinate and formate. The maximum growth rate, μ m, half saturation coefficients, K s, microbial yield coefficient, Y, cell mass decay rate coefficient, k d, and substrate inhibition coefficient, K si were 0.25 ± 0.03 h−1, 118 ± 31 mg glucose l−1, 0.12 μg DNA mg glucose−1, 0.01 h−1, and 3,108 ± 1,152 mg glucose l−1, respectively. These data were used to predict the performance of a continuous growth system with an influent glucose concentration of 5,000 mg l−1. Results of the analysis suggest that A. hydrophila will metabolize glucose at greater than 95% efficiency when hydraulic retention times (HRTs) exceed 7 h, whereas the culture is at risk of washing out at an HRT of 6.7 h.  相似文献   

12.
Formate oxidase was found in cell-free extracts of Debaryomyces vanrijiae MH201, a soil isolate. After purification by column chromatography, the preparation showed a protein band corresponding to a molecular mass (MM) of 64 kDa on sodium dodecyl sulfate–polyacrylamide gel electrophoresis. The MM, estimated by a gel filtration, was 99 kDa. The preparation showed two and three bands on isoelectric focusing under denaturing and native conditions, respectively. These results suggest that the preparation contained three isoforms, each of which might be composed of αα, αβ, and ββ subunits with apparently similar MM. The preparation acted on formate with K m and V max values of 11.7 mM and 262 μmol min−1 mg−1, respectively, at pH 4.5 and 25°C, but showed no evidence of activity on the other compounds tested. The optimum pH and temperature were pH 4.0 and 35°C, respectively. The preparation showed activities of 85% of the initial activity after storage at pH 6.0 and 4°C for 8 weeks. When 10 mM formaldehyde was reacted with 2.0 U ml−1 of the enzyme preparation at pH 5.5 and room temperature in the presence of 2.0 U ml−1 of a microbial aldehyde oxidase and 100 U ml−1 of catalase for 180 min, neither of formate nor formaldehyde was detected, suggesting that the reaction involved the quantitative conversion of formaldehyde to carbon dioxide.  相似文献   

13.
To investigate annual variation in soil respiration (R S) and its components [autotrophic (R A) and heterotrophic (R H)] in relation to seasonal changes in soil temperature (ST) and soil water content (SWC) in an Abies holophylla stand (stand A) and a Quercus-dominated stand (stand Q), we set up trenched plots and measured R S, ST and SWC for 2 years. The mean annual rate of R S was 436 mg CO2 m−2 h−1, ranging from 76 to 1,170 mg CO2 m−2 h−1, in stand A and 376 mg CO2 m−2 h−1, ranging from 82 to 1,133 mg CO2 m−2 h−1, in stand Q. A significant relationship between R S and its components and ST was observed over the 2 years in both stands, whereas a significant correlation between R A and SWC was detected only in stand Q. On average over the 2 years, R A accounted for approximately 34% (range 17–67%) and 31% (15–82%) of the variation in R S in stands A and Q, respectively. Our results suggested that vegetation type did not significantly affect the annual mean contributions of R A or R H, but did affect the pattern of seasonal change in the contribution of R A to R S.  相似文献   

14.
 Random amplified polymorphic DNA (RAPD) markers were identified for self-incompatibility (SI) alleles that will allow marker-assisted selection of desired S-alleles in hazelnut (Corylus avellana L.). DNA was extracted from young leaves collected from field-planted parents and 26 progeny of the cross OSU 23.017 (S1S12)×VR6-28 (S2S26) (OSU23×VR6). Screening of 10-base oligonucleotide RAPD primers was performed using bulked segregant analysis. DNA samples from 6 trees each were pooled into four ‘bulks’, one for each of the following: S1 S2, S1 S26 , S2 S12, and S12 S26. ‘Super bulks’ of 12 trees each for S1, S2, S12, and S26 were then created for each allele by combining the appropriate bulks. The DNA from these four super bulks and from the parents was used as a template in the PCR assays. A total of 250 primers were screened, and one RAPD marker each was identified for alleles S2 (OPI07750) and S1 (OPJ141700). OPJ141700 was identified in 13 of 14 S1 individuals of the cross OSU23×VR6 used in bulking and yielded a false positive in 1 non-S1 individual. This same marker was not effective outside the original cross, identifying 4 of 5 S1 progeny in another cross, ‘Willamette’×VR6-28 (‘Will’×VR6), but yielded false positives in 4 of 9 non-S1 individuals from the cross ‘Casina’×VR6-28 (‘Cas’×VR6). OPI07750 served as an excellent marker for the S2 allele and was linked closely to this allele, identifying 12 of 13 S2 individuals in the OSU23×VR6 population with no false positives. OPI07750 was found in 4 of 4 S2 individuals from ‘Will’×VR and 7 of 7 S2 individuals of ‘Cas’×VR6 with no false positives, as well as 10 of 10 S2 individuals of the cross OSU 296.082 (S1S8)×VR8-32 (S2S26), with only 1 false positive individual out of 21 progeny. OPI07750 was also present in 5 of 5 cultivars carrying the S2 allele, with no false-positive bands in non-S2 cultivars, and correctly identified all but 2 S2 individuals in 57 additional selections in the breeding program. In the OSU23×VR6 population, the recombination rate between the marker OPJ141700 and the S1 allele was 7.6% and between the OPI07750 marker and the S2 allele was 3.8%. RAPD marker bands were excised from gels, cloned, and sequenced to enable the production of longer primers (18 or 24 bp) that were used to obtain sequence characterized amplified regions (SCARs). Both the S1 and S2 markers were successfully cloned and 18 bp primers yielded the sole OPJ141700 product, while 24-bp primers yielded OPI07750 as well as an additional smaller product (700 bp) that was not polymorphic but was present in all of the S-genotypes examined. Received: 10 January 1998 / Accepted: 26 January 1998  相似文献   

15.
Progressive microwave power saturation (P1/2) measurements have been performed on the tyrosine D radical (YD ) of photosystem II (PSII) in order to examine its relaxation enhancement by the oxygen-evolving complex (OEC) poised to the reduced S−1 and S−2 oxidation states by NO treatment. Analysis of the power saturation curves showed that the S−1 oxidation state of the OEC does not enhance the relaxation of YD : it therefore possesses a diamagnetic ground state. In contrast, the Mn(II)-Mn(III) multiline electron paramagnetic resonance (EPR) signal characteristic of the S−2 oxidation state of the OEC was shown to provide a relaxation enhancement pathway for YD , however less efficient relative to the one provided by the S2-state multiline EPR signal. We also examined the YD relaxation enhancement characteristics of the EPR-silent oxidation state produced after brief (1–5 min) dark incubation at 0°C of a PSII sample poised to the EPRactive S−2 state. This EPR-silent oxidation state denoted as “0°C incubation” state was shown to possess remarkably similar P1/2 values with the EPR-active S−2 state in the overall examined temperature range (6–20 K). In addition, these values remained unchanged after successive cycles of the OEC between the EPR-active S−2 state and the “0°C incubation” state. The data presented in this work point to the conclusion that the “0°C incubation” state is indeed an S−2 oxidation state with half-integer spin.  相似文献   

16.
The effect of trace metal ions (Co2+, Cu2+, Fe2+, Mn2+, Mo6+, Ni2+, Zn2+, SeO4 and WO4 ) on growth and ethanol production by an ethanologenic acetogen, Clostridium ragsdalei was investigated in CO:CO2-grown cells. A standard acetogen medium (ATCC medium no. 1754) was manipulated by varying the concentrations of trace metals in the media. Increasing the individual concentrations of Ni2+, Zn2+, SeO4 and WO4 from 0.84, 6.96, 1.06, and 0.68 μM in the standard trace metals solution to 8.4, 34.8, 5.3, and 6.8 μM, respectively, increased ethanol production from 35.73 mM under standard metals concentration to 176.5, 187.8, 54.4, and 72.3 mM, respectively. Nickel was necessary for growth of C. ragsdalei. Growth rate (μ) of C. ragsdalei improved from 0.34 to 0.49 (day−1), and carbon monoxide dehydrogenase (CODH) and hydrogenase (H2ase)-specific activities improved from 38.45 and 0.35 to 48.5 and 1.66 U/mg protein, respectively, at optimum concentration of Ni2+. At optimum concentrations of WO4 and SeO4 , formate dehydrogenase (FDH) activity improved from 32.3 to 42.6 and 45.4 U/mg protein, respectively. Ethanol production and the activity of FDH reduced from 35 mM and 32.3 U/mg protein to 1.14 mM and 8.79 U/mg protein, respectively, upon elimination of WO4 from the medium. Although increased concentration of Zn2+ enhanced growth and ethanol production, the activities of CODH, FDH, H2ase and alcohol dehydrogenase (ADH) were not affected by varying the Zn2+ concentration. Omitting Fe2+ from the medium decreased ethanol production from 35.7 to 6.30 mM and decreased activities of CODH, FDH, H2ase and ADH from 38.5, 32.3, 0.35, and 0.68 U/mg protein to 9.07, 7.01, 0.10, and 0.24 U/mg protein, respectively. Ethanol production improved from 35 to 54 mM when Cu2+ was removed from the medium. The optimization of trace metals concentration in the fermentation medium improved enzyme activities (CODH, FDH, and H2ase), growth and ethanol production by C. ragsdalei.  相似文献   

17.
The microbial population of geothermally heated sediments in a shallow bay of Vulcano Island (Italy) was characterized with respect to metabolic activities and the putatively catalyzing hyperthermophiles. Site-specific anoxic culturing media, most of which were amended with combinations of electron donors (glucose or carboxylic acids) and acceptors (sulfate), were used for selective enrichment of metabolically defined subpopulations. The mostly archaeal chemoautotrophs produced formate at rates of 3.25 and 0.46 fmol cell−1 day−1 with and without sulfate, respectively. The glucose fermenting heterotrophs produced acetate (18 fmol cell−1 day−1) and lactate (2.6 fmol cell−1 day−1) and were identified as predominantly Thermus sp. and coccoid archaea. These archaeal cells also metabolized lactate (5.6 fmol cell−1 day−1), but neither formate nor acetate. The heterotrophic culture enriched on formate/acetate/propionate/sulfate utilized mainly formate (27 fmol cell−1 day−1) and lactate (89–195 fmol cell−1 day−1), and consumed sulfate (38–68 fmol cell−1 day−1). These formate or lactate consuming sulfate reducers were dominated by Archaeoglobales (7% in situ) and unidentified Archaea. The in situ benthic community comprised 15% Crenarchaeota, a significant group only in the autotrophic cultures, and 3% Thermus sp., the putatively predominant group involved in fermentative metabolism. The role of Thermoccales (4% in situ) remained undisclosed in our experiments. This first comprehensive data set established plausible links between several groups of hyperthermophiles in shallow marine hydrothermal systems, their metabolic function within the benthic microbial community, and biogeochemical turnover rates.  相似文献   

18.
Electrospray ionisation (ESI) mass spectrometry was used to examine the reactions of the clinically used antiarthritic agent [Au(S2O3)2]3−, and AuPEt3Cl, a derivative of another clinically used agent auranofin, with human serum albumin (HSA) obtained from a human volunteer. Both compounds reacted readily with HSA to form complexes containing one or more covalently attached gold fragments. In the case of AuPEt3Cl, binding was accompanied by the loss of the chloride ligand, while for [Au(S2O3)2]3− the mass spectral data indicated binding of Au(S2O3) groups. Experiments performed using HSA with Cys34 blocked by reaction with iodoacetamide were consistent with reaction of both gold compounds with this amino acid. Separate blocking experiments using diethylpyrocarbonate and AuPEt3Cl also provided evidence for histidine residues acting as lower-affinity binding sites for this gold compound. ESI mass spectra of solutions containing [Au(S2O3)2]3− or [Au(CN)2], and HSA, provided evidence for the formation of protein complexes in which intact gold molecules were non-covalently bound. In the case of [Au(S2O3)2]3−, these non-covalent complexes proved to be transitory in nature. However, for [Au(CN)2] a non-covalent complex containing a single gold molecule bound to HSA was found to be stable, and constituted the main adduct formed in solutions containing low-to-medium Au-to-HSA ratios. Evidence was also obtained for the formation of a covalent adduct in which a single Au(CN) moiety was bonded to Cys34 of the protein. AuPEt3Cl reacted to a much lower extent with HSA that had Cys34 modified by formation of a disulfide bond to added cysteine, than with unmodified HSA. This suggests that the extent of modification of the protein in vivo may have an important influence on the transport and bioavailability of gold antiarthritic drugs.Electronic Supplementary Material Supplementary material is available for this article at and is accessible for authorized users.  相似文献   

19.
pH affected significantly the growth and the glucose fermentation pattern of Propionibacterium microaerophilum. In neutral conditions (pH 6.5–7.5), growth and glucose fermentation rate (qs) were optimum producing propionate, acetate, CO2, and formate [which together represented 90% (wt/wt) of the end products], and lactate representing only 10% (wt/wt) of the end products. In acidic conditions, propionate, acetate, and CO2 represented nearly 100% (wt/wt) of the fermentation end products, whereas in alkaline conditions, a shift of glucose catabolism toward formate and lactate was observed, lactate representing 50% (wt/wt) of the fermentation end products. The energy cellular yields (Y X/ATP), calculated (i) by taking into account extra ATP synthesized through the reduction of fumarate into succinate, was 6.1–7.2 g mol−1. When this extra ATP was omitted, it was 11.9–13.1 g mol−1. The comparison of these values with those of Y X/ATP in P. acidipropionici and other anaerobic bacteria suggested that P. microaerophilum could not synthesize ATP through the reduction of fumarate into succinate and therefore differed metabolically from P. acidipropionici. Received: 8 April 2002 / Accepted: 8 May 2002  相似文献   

20.
Immobilized cells of Delftia tsuruhatensis CCTCC M 205114 harboring R-amidase were applied in asymmetric hydrolysis of (R)-2, 2-dimethylcyclopropane carboxamide (R − 1) from racemic (R, S)-2, 2-dimethylcyclopropane carboxamide to accumulate (S)-2, 2-dimethylcyclopropane carboxamide (S − 1). Maximum R-amidase activity of 13.1 U/g wet cells (0.982 U/g beads) was obtained under conditions of 3% sodium alginate, 2.5% CaCl2, 15 h crosslinking and 2 mm bead size, which was 53.9% of that of free cells (24.3 U/g wet cells). In addition, characterization of the immobilized cells was examined. The optimum R − 1 hydrolysis conditions were identified as follows: substrate concentration 10 mM, pH 8.5, temperature 35°C and time course 40 min. Under optimum conditions, the maximum yield and enantiomeric excess of (R)-2, 2-dimethylcyclopropanecarboxylic acid were 49.5% and >99%, respectively. This afforded S − 1 with a yield >49% and an e.e. of 97.7%. With good operational stability and excellent enanotioselectivity, the immobilized cells could be potentially utilized in industrial production of S − 1.  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