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.
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   

2.
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   

3.
Copper and other transition metal ions and their complexes are catalysts for the decomposition of nitrosothiols. In this way they catalyze the biological functions of nitrosothiols. The kinetics and mechanism of the reaction of two nitrosothiols, S-nitrosothiolactic acid and S-nitrosoglutathione (GSNO), with copper(I) are reported. The kinetics of the reaction of Cu(MeCN) n + (n=0–3) with the nitrosothiols were studied. The results indicate that Cu+ aq is the active species in the GSNO system, with k(Cu+ aq+GSNO)=(9.4 ±2.0)×107 dm3 mol−1 s−1 . The results also indicate that the Cu(MeCN) n + (n=0–3) complexes react with S-nitrosothiolactic acid. Transient species are formed in these processes. The results suggest that these species contain copper(I) and thiol. The results shed light on the catalytic role of copper complexes in the decomposition of S-nitrosothiols. Received 10 April 1999 / Accepted 17 December 1999  相似文献   

4.
Molybdenum and tungsten complexes as models for the active sites of assimilatory or dissimilatory nitrate reductases (NR) were computed at the CPCM-B98/SDDp//B3LYP/Lanl2DZp* plus zero point energy level of density functional theory. The ligands were chosen on the basis of available experimental protein or small chemical model structures. A water molecule is found to bind to assimilatory NR models [(Me2C2S2)MO(YMe)] (−11.5 kcal mol−1 for M is Mo, Y is S) and may be replaced by nitrate (−4.5 kcal mol−1) (but a hydroxy group may not). Nature’s choice of M is Mo and Y is S for NR has the largest activation energy for protein-free models (13.3 kcal mol−1) and the least exothermic reaction energy for the nitrate reduction (−14.9 kcal mol−1) compared with M is W and Y is O or Se alternatives. Water binding to dissimilatory NR model complexes [(Me2C2S2)2M(YR)] is considerably endothermic (10.3 kcal mol−1); nitrate binding is only slightly so (1.5 kcal mol−1 for RY is MeS). The exchange of an oxo ligand (assimilatory NR) for a dithiolato ligand (dissimilatory NR model) reduces the exothermicity (−8.6 kcal mol−1 relative to the fivefold-coordinate reduced complex) and raises the barrier for oxygen atom transfer (OAT) in the nitrate complex (19.2 kcal mol−1). Not for the mono but only for the bisdithiolato complexes hydrogen bonding involving the coordinated substrate may significantly lower the OAT barrier as shown by explicitly adding water molecules. Substitution of tungsten for molybdenum generally lowers OAT activation energies and makes nitrate reduction reaction energies more negative. Bidentate carboxylato binding identified in Escherichia coli NarGHI is the preferred binding mode also for an acetato model. However, one dithiolato ligand folds when the MoVI center is bare of a good π-donor ligand, e.g., an oxo group. Computations on [(mnt)2MoIV(YR)(PPh3)] [mnt is (CN)2C2S2 2−] gave a smaller nitrate reduction activation energy for RY is Cl, compared with RY is PhS, although experimentally only the phenyl thiolato complex and not the chloro complex was found to be a functional NR model. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

5.
The cyanide-degrading fungusRhizopus oryzae associated with post-harvest spoilage of cassava (Manihot esculenta L.) oxidized S0 to S2O3 2−, S4O6 2− and SO4 2− in culture and when grown in autoclaved soil amended with S0. Oxidation of sulfur was associated with rhodanese activity of the fungus.  相似文献   

6.
In order to understand the detailed mechanism of the stereoselective photoinduced electron-transfer (ET) reactions of zinc-substituted myoglobin (ZnMb) with optically active molecules by flash photolysis, we designed and prepared new optically active agents, such as N,N′-dimethylcinchoninium diiodide ([MCN]I2) and N,N′-dimethylcinchonidinium diiodide ([MCD]I2). The photoexcited triplet state of ZnMb, 3(ZnMb)*, was successfully quenched by [MCN]2+ and [MCD]2+ ions to form the radical pair of ZnMb cation (ZnMb·+) and reduced [MCN]·+ and [MCD]·+, followed by a thermal back ET reaction to the ground state. The rate constants (k q) for the ET quenching at 25 °C were obtained as k q(MCN)=(1.9±0.1)×106 M−1 s−1 and k q(MCD)=(3.0±0.2)×106 M−1 s−1, respectively. The ratio of k q(MCD)/k q(MCN)=1.6 indicates that the [MCD]2+ preferentially quenches 3(ZnMb)*. The second-order rate constants (k b) for the thermal back ET reaction from [MCN]·+ and [MCD]·+ to ZnMb·+ at 25 °C were k b(MCN)=(0.79±0.04)×108 M−1 s−1 and k b(MCD)=(1.0±0.1)×108 M−1 s−1, respectively, and the selectivity was k q(MCD)/k q(MCN)=1.3. Both quenching and thermal back ET reactions are controlled by the ET step. In the quenching reaction, the energy differences of ΔΔH (MCD–MCN) and ΔΔS (MCD–MCN) at 25 °C were obtained as −1.1 and 0 kJ mol−1, respectively. On the other hand, ΔΔH (MCD–MCN)=11±2 kJ mol−1 and TΔΔS (MCD–MCN)=−10±2 kJ mol−1 were given in the thermal back ET reaction. The highest stereoselectivity of 1.7 for [MCD]·+ found at low temperature (10 °C) was due to the ΔΔS value obtained in the thermal back ET reaction. Electronic Supplementary Material Supplementary material is available for this article at and is accessible for authorized users.  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.
Our basic knowledge of the ecology, especially the age and growth of polar deep-sea biota is still scarce. This study provides first data about the age and growth of the two abundant Arctic fish species Lycodes frigidus and Lycodes squamiventer (Zoarcidae). Lycodes frigidus was caught at the deeper parts (1,546–3,576 m depth) of the HAUSGARTEN observatory (HG), west of Svalbard. The congener Lycodes squamiventer was caught at two HG stations (1,273–1,546 m) and at the H?kon Mosby Mud Volcano (HMMV, ~1,250 m), a cold seep in the southwestern Barents Sea. Age was determined by sagittal otolith increment analysis. Growth performance was assessed by fitting age–length data to a von Bertalanffy growth equation. Our data suggest that L. frigidus and L. squamiventer attain maximum ages of 33 and 21 years, respectively. Lycodes squamiventer from the HMMV had significantly higher growth rates and their maximum age and length was slightly lower compared to conspecifics from the shallow HG stations. Von Bertalanffy growth equations were L t  = 58.9 ∗ (1 − e(−0.042*t)) for L. frigidus, and L t  = 25.3 ∗ (1 − e(−0.074*t)) and L t  = 24.2 ∗ (1 − e(−0.099 * t)) for L. squamiventer from HG and the HMMV, respectively. A comparison of these data with those of eight other zoarcids indicates that growth performances are correlated with temperature: the higher the annual mean temperatures experienced, the higher the growth rates. However, maximum ages decrease with increasing temperatures.  相似文献   

9.
The occurrence of biochemical activities of the sulphur cycle was followed in isolates of heterotrophic bacteria from the fermentative horizon of a spruce stand, a grass-covered withered spruce stand and of mountain ash and birch stand in the area strongly influenced by sulphur immissions. The occurrence of bacteria capable of reducing S0 to S2−, oxidizing S0 and S2O3 2− to SO4 2− and solubilizing S0 increased in the above order. The occurrence of producers of thiosulphate sulphurtransferase (rhodanese), thiosulphate oxidase and sulphite oxidase increased and the level of the production of these enzymes increased as well. Heterotrophic bacteria (mostly pseudomonads) from the grass-covered stands exhibit more activities of the sulphur cycle than bacteria from the spruce stand without ground vegetation.  相似文献   

10.
The Ferrous Wheel Hypothesis (Davidson et al. 2003) postulates the abiotic formation of dissolved organic N (DON) in forest floors, by the fast reaction of NO2 with dissolved organic C (DOC). We investigated the abiotic reaction of NO2 with dissolved organic matter extracted from six different forest floors under oxic conditions. Solutions differed in DOC concentrations (15–60 mg L−1), NO2 concentrations (0, 2, 20 mg NO2 -N L−1) and DOC/DON ratio (13.4–25.4). Concentrations of added NO2 never decreased within 60 min, therefore, no DON formation from added NO2 took place in any of the samples. Our results suggest that the reaction of NO2 with natural DOC in forest floors is rather unlikely.  相似文献   

11.
This article reports rate constants for thiol–thioester exchange (k ex), and for acid-mediated (k a), base-mediated (k b), and pH-independent (k w) hydrolysis of S-methyl thioacetate and S-phenyl 5-dimethylamino-5-oxo-thiopentanoate—model alkyl and aryl thioalkanoates, respectively—in water. Reactions such as thiol–thioester exchange or aminolysis could have generated molecular complexity on early Earth, but for thioesters to have played important roles in the origin of life, constructive reactions would have needed to compete effectively with hydrolysis under prebiotic conditions. Knowledge of the kinetics of competition between exchange and hydrolysis is also useful in the optimization of systems where exchange is used in applications such as self-assembly or reversible binding. For the alkyl thioester S-methyl thioacetate, which has been synthesized in simulated prebiotic hydrothermal vents, k a = 1.5 × 10−5 M−1 s−1, k b = 1.6 × 10−1 M−1 s−1, and k w = 3.6 × 10−8 s−1. At pH 7 and 23°C, the half-life for hydrolysis is 155 days. The second-order rate constant for thiol–thioester exchange between S-methyl thioacetate and 2-sulfonatoethanethiolate is k ex = 1.7 M−1 s−1. At pH 7 and 23°C, with [R″S(H)] = 1 mM, the half-life of the exchange reaction is 38 h. These results confirm that conditions (pH, temperature, pK a of the thiol) exist where prebiotically relevant thioesters can survive hydrolysis in water for long periods of time and rates of thiol–thioester exchange exceed those of hydrolysis by several orders of magnitude.  相似文献   

12.
Both HA-CdS and HB-CdS (Hys-CdS, Hys represents HA, HB) complex systems were established according to the dynamics of heterogeneous electron-transfer process m = ES* /S+ - E < 0\mu = E_{S^* /S^ + } - E< 0 . In these systems, the electron transferring from1Hys* to conduction band of CdS is feasible. Determined from the fluorescence quenching, the apparent association constants (Kapp) between Hypocrellin A (HA), Hypocrellin B. (HB) and CdS sol. were about 940 (mol/L)−1, 934 (mol/L)−1, respectively. Fluorescence lifetime measurements gave the rate constant for the electron transfer process from1HA*,1HB* into conduction band of CdS semiconductor as 5.16 × 109 s−1, 5.10 × 109 s−1, respectively. TEMPO (2,2,6,6-tetramethy-1-piperdinyloxy), a stable nitroxide radical, was used in the kinetic study of the reduction reaction taking place on the surface of a CdS colloidal semiconductor, kinetics equation of the reaction was determined with the electron paramagnetic resonance (EPR) method, and the reaction order of TEMPO is zero. When Hys were added, the rate of EPR increased greatly. By comparing rate constants, the Hys-CdS systems were revealed to be about 350 times more efficient than CdS sol. alone in the photoreduction of TEMPO under visible light. It suggests that Hys can be used as efficient sensitizers of a colloidal semiconductor in the application of solar energy.  相似文献   

13.
Synopsis The routine swimming speed (S) of three groups of 4, 9 and 32 cm total length (LT) juvenile cod (Gadus morhua) was quantified in the laboratory at 6 – 10 different temperatures (T) between 3.2 and 16.7°C. At temperatures between 5 and 15°C, mean group S increased exponentially with increasing T (S=a ebT) and the effect of temperature (b = 0.082, Q10 = 2.27) was not significantly different among the groups (over the 8-fold difference in fish sizes of early- and post-settlement juveniles). Differences in mean S among individuals within each group were quite large (coefficient of variation = 40 – 80%). Swimming data for juveniles and those collected for groups of 0.4, 0.7 and 0.9 cm standard length (LS) larvae were combined to assess the effect of body size on S. At 8°C, S (mm s−1) increased with LS (mm) according to: S = 0.26LSΦ−5.28LS−1, where Φ = 1.55LS−0.08. Relative S (body lengths s−1) was related to LS by a dome-shaped relationship having a maximum value (0.49 body lengths s−1) at 18.5 – 19 mm LS corresponding to the sizes of fish at the end of larval-juvenile metamorphosis. Previous larval cod IBM’s using a cruise-predator mode likely overestimated rates of foraging (prey searching and encounters) by a factor of ~2, whereas foraging rates in pause-travel models are closer to estimates of swimming velocities obtained in this and other laboratory studies.  相似文献   

14.
Reservoirs are intrinsically linked to the rivers that feed them, creating a river–reservoir continuum in which water and sediment inputs are a function of the surrounding watershed land use. We examined the spatial and temporal variability of sediment denitrification rates by sampling longitudinally along an agriculturally influenced river–reservoir continuum monthly for 13 months. Sediment denitrification rates ranged from 0 to 63 μg N2O g ash free dry mass of sediments (AFDM)−1 h−1 or 0–2.7 μg N2O g dry mass of sediments (DM)−1 h−1 at reservoir sites, vs. 0–12 μg N2O gAFDM−1 h−1 or 0–0.27 μg N2O gDM−1 h−1 at riverine sites. Temporally, highest denitrification activity traveled through the reservoir from upper reservoir sites to the dam, following the load of high nitrate (NO3-N) water associated with spring runoff. Annual mean sediment denitrification rates at different reservoir sites were consistently higher than at riverine sites, yet significant relationships among theses sites differed when denitrification rates were expressed per gDM vs. per gAFDM. There was a significant positive relationship between sediment denitrification rates and NO3-N concentration up to a threshold of 0.88 mg NO3 -N l−1, above which it appeared NO3-N was no longer limiting. Denitrification assays were amended seasonally with NO3-N and an organic carbon source (glucose) to determine nutrient limitation of sediment denitrification. While organic carbon never limited sediment denitrification, all sites were significantly limited by NO3-N during fall and winter when ambient NO 3-N was low.  相似文献   

15.
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   

16.
We investigated the influence of stand density [938 tree ha−1 for high stand density (HD), 600 tree ha−1 for medium stand density (MD), and 375 tree ha−1 for low stand density (LD)] on soil CO2 efflux (R S) in a 70-year-old natural Pinus densiflora S. et Z. forest in central Korea. Concurrent with R S measurements, we measured litterfall, total belowground carbon allocation (TBCA), leaf area index (LAI), soil temperature (ST), soil water content (SWC), and soil nitrogen (N) concentration over a 2-year period. The R S (t C ha−1 year−1) and leaf litterfall (t C ha−1 year−1) values varied with stand density: 6.21 and 2.03 for HD, 7.45 and 2.37 for MD, and 6.96 and 2.23 for LD, respectively. In addition, R S was correlated with ST (R 2 = 0.77–0.80, P < 0.001) and SWC (R 2 = 0.31–0.35, P < 0.001). It appeared that stand density influenced R S via changes in leaf litterfall, LAI and SWC. Leaf litterfall (R 2 = 0.71), TBCA (R 2 = 0.64–0.87), and total soil N contents in 2007 (R 2 = 0.94) explained a significant amount of the variance in R S (P < 0.01). The current study showed that stand density is one of the key factors influencing R S due to the changing biophysical and environmental factors in P. densiflora.  相似文献   

17.
Inhibition of electron transport and damage to the protein subunits by visible light has been studied in isolated reaction centers of the non-sulfur purple bacterium Rhodobacter sphaeroides. Illumination by 1100 μEm−2 s−1 light induced only a slight effect in wild type, carotenoid containing 2.4.1. reaction centers. In contrast, illumination of reaction centers isolated from the carotenoidless R26 strain resulted in the inhibition of charge separation as detected by the loss of the initial amplitude of absorbance change at 430 nm arising from the P+QB → PQB recombination. In addition to this effect, the L, M and H protein subunits of the R26 reaction center were damaged as shown by their loss on Coomassie stained gels, which was however not accompanied by specific degradation products. Both the loss of photochemical activity and of protein subunits were suppressed in the absence of oxygen. By applying EPR spin trapping with 2,2,6,6-tetramethylpiperidine we could detect light-induced generation of singlet oxygen in the R26, but not in the 2.4.1. reaction centers. Moreover, artificial generation of singlet oxygen, also led to the loss of the L, M and H subunits. Our results provide evidence for the common hypothesis that strong illumination by visible light damages the carotenoidless reaction center via formation of singlet oxygen. This mechanism most likely proceeds through the interaction of the triplet state of reaction center chlorophyll with the ground state triplet oxygen in a similar way as occurs in Photosystem II. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

18.
Hydrofluoroethers are being considered as potential candidates for third generation refrigerants. The present investigation involves the ab initio quantum mechanical study of the decomposition mechanism of CF3OCH2O radical formed from a hydrofluoroether, CF3OCH3 (HFE-143a) in the atmosphere. The geometries of the reactant, products and transition states involved in the decomposition pathways are optimized and characterized at the DFT (B3LYP) level of theory using 6-311G(d,p) basis set. Energy calculations have been performed at the G2(MP2) and G2M(CC,MP2) level of theory. Two prominent decomposition channels, C-O bond scission and reaction with atmospheric O2 have been considered for detailed investigation. Studies performed at the G2(MP2) level reveals that the decomposition channel involving C-O bond scission occurs with a barrier height of 23.8 kcal mol−1 whereas the oxidative pathway occurring with O2 proceeds with an energy barrier of 7.2 kcal mol−1. On the other hand the corresponding values at G2M(CC,MP2) are 24.5 and 5.9 kcal mol−1 respectively. Using canonical transition state theory (CTST) rate constants for the two pathways considered are calculated at 298 K and 1 atm pressure and found to be 5.9 × 10−6 s−1 and 2.3 × 10−5 s−1 respectively. The present study concludes that reaction with O2 is the dominant path for the consumption of CF3OCH2O in the atmosphere. Transition states are searched and characterized on the potential energy surfaces involved in both of the reaction channels. The existence of transition state on the corresponding potential energy surface is ascertained by performing intrinsic reaction coordinate (IRC) calculation.  相似文献   

19.
Kinetics of kojic acid fermentation by Aspergillus flavus Link 44-1 using various sources of carbon [glucose, xylose, sucrose, starch, maltose, lactose or fructose] and nitrogen [NH4Cl, (NH4)2S2O8, (NH4)2NO3, yeast extract or peptone] were analyzed using models based on logistic and Luedeking–Piret equations. The highest kojic acid production (39.90 g l−1) in submerged batch fermentation was obtained when 100 g l−1 glucose was used as a carbon source. Organic nitrogen sources such as peptone and yeast extract were favorable for kojic acid production as compared to inorganic nitrogen sources. Yeast extract at 5 g l−1 was optimal. The optimal carbon to nitrogen (C/N) ratio for kojic acid fermentation was 93.3. In a resuspended cell system, the rate of glucose conversion to kojic acid by cell-bound enzymes increased with increasing glucose concentration up to 70 g l−1, suggesting that the reaction followed the Michaelis–Menten enzyme kinetic model. The value of K m and V max for the reaction was 18.47 g l−1 glucose and 0.154 g l−1 h−1, respectively. Journal of Industrial Microbiology & Biotechnology (2000) 25, 20–24. Received 13 October 1999/ Accepted in revised form 02 April 2000  相似文献   

20.
The influence of reduced sulfur compounds (including stored S0) on H2 evolution/consumption reactions in the purple sulfur bacterium, Thiocapsa roseopersicina BBS, was studied using mutants containing only one of the three known [NiFe] hydrogenase enzymes: Hox, Hup or Hyn. The observed effects depended on the kind of hydrogenase involved. The mutant harbouring Hox hydrogenase was able to use S2O32−, SO32−, S2− and S0 as electron donors for light-dependent H2 production. Dark H2 evolution from organic substrates via Hox hydrogenase was inhibited by S0. Under light conditions, endogenous H2 uptake by Hox or Hup hydrogenases was suppressed by S compounds. СО2-dependent H2 uptake by Hox hydrogenase in the light required the additional presence of S compounds, unlike the Hup-mediated process. Dark H2 consumption via Hyn hydrogenase was connected to utilization of S0 as an electron acceptor and resulted in the accumulation of H2S. In wild type BBS, with high levels of stored S0, dark H2 production from organic substrates was significantly lower, but H2S accumulation significantly higher, than in the mutant GB1121(Hox+). There is a possibility that H2 produced via Hox hydrogenase is consumed by Hyn hydrogenase to reduce S0.  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