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.
NADH:ubiquinone oxidoreductase (Complex I), the electron input enzyme in the respiratory chain of mitochondria and many bacteria, couples electron transport to proton translocation across the membrane. Complex I is a primary proton pump; although its proton translocation mechanism is yet to be known, it is considered radically different from any other mechanism known for redox-driven proton pumps: no redox centers have been found in its membrane domain where the proton translocation takes place. Here we studied the properties and the catalytic role of the enzyme-bound ubiquinone in the solubilized, purified Complex I from Escherichia coli. The ubiquinone content in the enzyme preparations was 1.3±0.1 per bound FMN residue. Rapid mixing of Complex I with NADH, traced optically, demonstrated that both reduction and re-oxidation kinetics of ubiquinone coincide with the respective kinetics of the majority of Fe-S clusters, indicating kinetic competence of the detected ubiquinone. Optical spectroelectrochemical redox titration of Complex I followed at 270-280nm, where the redox changes of ubiquinone contribute, did not reveal any transition within the redox potential range typical for the membrane pool, or loosely bound ubiquinone (ca. +50-+100mV vs. NHE, pH 6.8). The transition is likely to take place at much lower potentials (E(m) ≤-200mV). Such perturbed redox properties of ubiquinone indicate that it is tightly bound to the enzyme's hydrophobic core. The possibility of two ubiquinone-binding sites in Complex I is discussed.  相似文献   

2.
Iron–sulfur cluster N2 of complex I (proton pumping NADH:quinone oxidoreductase) is the immediate electron donor to ubiquinone. At a distance of only ~ 7 Å in the 49-kDa subunit, a highly conserved tyrosine is found at the bottom of the previously characterized quinone binding pocket. To get insight into the function of this residue, we have exchanged it for six different amino acids in complex I from Yarrowia lipolytica. Mitochondrial membranes from all six mutants contained fully assembled complex I that exhibited very low dNADH:ubiquinone oxidoreductase activities with n-decylubiquinone. With the most conservative exchange Y144F, no alteration in the electron paramagnetic resonance spectra of complex I was detectable. Remarkably, high dNADH:ubiquinone oxidoreductase activities were observed with ubiquinones Q1 and Q2 that were coupled to proton pumping. Apparent Km values for Q1 and Q2 were markedly increased and we found pronounced resistance to the complex I inhibitors decyl-quinazoline-amine (DQA) and rotenone. We conclude that Y144 directly binds the head group of ubiquinone, most likely via a hydrogen bond between the aromatic hydroxyl and the ubiquinone carbonyl. This places the substrate in an ideal distance to its electron donor iron–sulfur cluster N2 for efficient electron transfer during the catalytic cycle of complex I.  相似文献   

3.
《Process Biochemistry》2014,49(5):740-744
The effects of oxido-reduction potential (ORP) control on succinic acid production have been investigated in Escherichia coli LL016. In LL016, two CO2 fixation pathways were achieved and NAD+ supply was enhanced by co-expression of heterologous pyruvate carboxylase (PYC) and nicotinic acid phosphoribosyltransferase (NAPRTase). During anaerobic fermentation, cell growth and metabolite distribution were changed with redox potential levels in the range of −200 to −400 mV. From the results, the ORP level of −400 mV was preferable, which resulted in the high succinic acid concentration (28.6 g/L) and high succinic acid productivity (0.33 g/L/h). Meanwhile, the yield of succinic acid at the ORP level of −400 mV was 39% higher than that at the ORP level of −200 mV. In addition, a higher NADH/NAD+ ratio and increased enzyme activities were also achieved by regulating the culture to a more reductive environment, which further enhanced the succinic acid production.  相似文献   

4.
《Process Biochemistry》2010,45(5):765-770
When Saccharomyces cerevisiae was grown under three glucose concentrations (ca. 200, 250, and 300 g/l), controlled at three reduction–oxidation (redox) potentials (no control, −150 and −100 mV) by manipulating two aerations (0.82 and 1.3 vvm), we observed that the recorded redox potential profiles resembled bathtub curves, and the profiles correlated well to the growth patterns measured under the same conditions. According to the shape of bathtub curve, we subdivided the curve into four regions. Region I features an abrupt decline in redox potential (corresponding to the growth phase from lag and logarithmic to the onset of stationary phase) that correlates to rapid yeast propagation, resulting from fast glucose uptake. Region II (corresponding to the stationary phase in yeast growth, characterizes a constant level of redox potential) is maintained by proper sparging and constant agitation. The continual buildup of ethanol causes growth arrest of yeast, resulting in the reduction of net NADH production. As a result, an uprising of redox potential is the feature of Region III, which signifies the end of stationary phase followed by the commencement of death phase. The severity of growth environment due to ethanol toxicity results in a rapid decrease in yeast population. Region IV (corresponding to the death phase during yeast growth) characterizes a drastic reduction in yeast viability and a gradual leveling of redox potential. A low glucose feed correlates to a fast decline of redox potential, a small basin in the bathtub curve, short fermentation duration, and complete glucose utilization. Imposing the current redox potential settings to low glucose feeds exerts no appreciable effect on ethanol production. In contrast, a high glucose feed connects to a sluggish bathtub curve for all four regions and incomplete glucose utilization. Proximate analysis on carbon balance indicates that controlling redox potential at −150 mV and under ca. 250 and 300 g glucose/l conditions, gave the highest fermentation efficiency as compared to other conditions; but there were no beneficiary effect to control redox potential under ca. 200 g glucose/l conditions.  相似文献   

5.
A combined DFT/electrostatic approach is employed to study the coupling of proton and electron transfer reactions in cytochrome c oxidase (CcO) and its proton pumping mechanism. The coupling of the chemical proton to the internal electron transfer within the binuclear center is examined for the O  E transition. The novel features of the His291 pumping model are proposed, which involve timely well-synchronized sequence of the proton-coupled electron transfer reactions. The obtained pKas and Ems of the key ionizable and redox-active groups at the different stages of the O  E transition are consistent with available experimental data. The PT step from E242 to H291 is examined in detail for various redox states of the hemes and various conformations of E242 side-chain. Redox potential calculations of the successive steps in the reaction cycle during the O  E transition are able to explain a cascade of equilibria between the different intermediate states and electron redistribution between the metal centers during the course of the catalytic activity. All four electrometric phases are discussed in the light of the obtained results, providing a robust support for the His291 model of proton pumping in CcO. This article is part of a Special Issue entitled: Respiratory oxidases.  相似文献   

6.
AtTDX is an enzyme present in Arabidopsis thaliana which is composed of two domains, a thioredoxin (Trx)-motif containing domain and a tetratricopeptide (TPR)-repeat domain. This enzyme has been shown to function as both a thioredoxin and a chaperone. The midpoint potential (Em) of AtTDX was determined by redox titrations using the thiol-specific modifiers, monobromobimane (mBBr) and mal-PEG. A NADPH/Trx reductase (NTR) system was used both to validate these Em determination methods and to demonstrate that AtTDX is an electron-accepting substrate for NTR. Titrations of full-length AtTDX revealed the presence of a single two-electron couple with an Em value of approximately ?260 mV at pH 7.0. The two cysteines present in a typical, conserved Trx active site (WCGPC), which are likely to play a role in the electron transfer processes catalyzed by AtTDX, have been replaced by serines by site-directed mutagenesis. These replacements (i.e., C304S, C307S, and C304S/C307S) resulted in a complete loss of the redox process detected using either the mBBr or mal-PEG method to monitor disulfide/dithiol redox couples. This result supports the conclusion that the couple with an Em value of ?260 mV is a disulfide/dithiol couple involving Cys304 and Cys307. Redox titrations for the separately-expressed Trx-motif containing C-domain also revealed the presence of a single two-electron couple with an Em value of approximately ?260 mV at 20 °C. The fact that these two Em values are identical, provides additional support for assignment of the redox couple to a disulfide/dithiol involving C304 and C307. It was found that, while the disulfide/dithiol redox chemistry of AtTDX was not affected by increasing the temperature to 40 °C, no redox transitions were observed at 50 °C and higher temperatures. In contrast, Escherichia coli thioredoxin was shown to remain redox-active at temperatures as high as 60 °C. The temperature-dependence of the AtTDX redox titration is similar to that observed for the redox activity of the protein in enzymatic assays.  相似文献   

7.
During mixed-acid fermentation by Corynebacterium crenatum under anaerobic conditions, two moles of NADH are required to synthesize 1 mol of succinic acid. In this work, four controlled culture redox potentials and different carbon sources with different oxidation states were used to investigate the possibility of enhancing the succinic acid production by increasing the availability of NADH. When the culture redox potential was ?300 mV, the yield of succinic acid was 0.31 g/g, representing a 72% increase compared with the yield when the culture redox potential was ?40 mV. Meanwhile, the molar ratio of succinic acid/lactic acid increased from 0.27 to 0.48. When 0.1% neutral red was added to the acid production medium, the yield of succinic acid was 0.25 g/g, and the molar ratio of succinic acid/lactic acid was 0.38. Both values were higher than those obtained from glucose only (0.19 g/g, 0.26) or gluconate (0.05 g/g, 0.18). A higher NADH/NAD+ ratio and increased enzymatic activity could be achieved to enhance the succinic acid production by manipulating the culture to a more reductive environment.  相似文献   

8.
Rearrangements of mitochondrial DNA in MSC16 mutant of cucumber (Cucumis sativus L.) affect mitochondrial functioning due to the alteration mainly of Complex I resulting in several metabolic changes. One-dimensional Blue-Native polyacrylamide gel electrophoresis (BN-PAGE) and densitometric measurements showed that the level and in-gel capacity of Complex I were lower in MSC16 leaf and root mitochondria as compared to wild-type (WT). The level and capacity of supercomplex I + III2 were always lower in leaf but not in MSC16 root mitochondria. Two-dimensional BN/SDS-PAGE indicated that the band abundance for most of the subunits of Complex I was lower in MSC16 leaf and root mitochondria. Supercomplex I + III2 level was only altered in MSC16 leaf mitochondria as measured after 2D BN/SDS-PAGE. No differences in the qualitative composition of the subunits of Complex I and supercomplex I + III2 between MSC16 and WT mitochondria were observed. In MSC16 mitochondria Complex I impairment could be compensated to some extent by additional respiratory chain NADH dehydrogenases. A higher capacity and level of NDB-1 protein of external NADH dehydrogenase was observed in MSC16 leaf and root mitochondria as compared to WT. The level of COX II, mitochondrial-encoded subunit of Complex IV, was higher in MSC16 leaf and root mitochondria. However, the capacity of Complex IV was slightly higher only in MSC16 leaf mitochondria. The levels of complexes: III2 and V and Complex V capacity did not differ in mitochondria between genotypes. An abundance of the subunits of respiratory complexes is one of the key factors determining not only their structure and functional stability but also a formation of the supercomplexes. We discuss here mitochondrial genome rearrangements in MSC16 mutant in a relation to assembly and/or stability (the lower level and capacity) of Complex I and supercomplex I + III2.  相似文献   

9.
In this paper allosteric interactions in protonmotive heme aa3 terminal oxidases of the respiratory chain are dealt with. The different lines of evidence supporting the key role of H+/e? coupling (redox Bohr effect) at the low spin heme a in the proton pump of the bovine oxidase are summarized. Results are presented showing that the I-R54M mutation in P. denitrificans aa3 oxidase, which decreases by more than 200 mV the Em of heme a, inhibits proton pumping. Mutational aminoacid replacement in proton channels, at the negative (N) side of membrane-inserted prokaryotic aa3 oxidases, as well as Zn2 + binding at this site in the bovine oxidase, uncouples proton pumping. This effect appears to result from alteration of the structural/functional device, closer to the positive, opposite (P) surface, which separates pumped protons from those consumed in the reduction of O2 to 2 H2O. This article is part of a Special Issue entitled: Respiratory Oxidases.  相似文献   

10.
Pyocyanin (N-methyl-1-hydroxyphenazine), a redox-active virulence factor produced by the human pathogen Pseudomonas aeruginosa, is known to compromise mucociliary clearance. Exposure of human bronchial epithelial cells to pyocyanin increased the rate of cellular release of H2O2 threefold above the endogenous H2O2 production. Real-time measurements of the redox potential of the cytosolic compartment using the redox sensor roGFP1 showed that pyocyanin (100 μM) oxidized the cytosol from a resting value of − 318 ± 5 mV by 48.0 ± 4.6 mV within 2 h; a comparable oxidation was induced by 100 μM H2O2. Whereas resting Cl secretion was slightly activated by pyocyanin (to 10% of maximal currents), forskolin-stimulated Cl secretion was inhibited by 86%. The decline was linearly related to the cytosolic redox potential (1.8% inhibition/mV oxidation). Cystic fibrosis bronchial epithelial cells homozygous for ΔF508 CFTR failed to secrete Cl in response to pyocyanin or H2O2, indicating that these oxidants specifically target the CFTR and not other Cl conductances. Treatment with pyocyanin also decreased total cellular glutathione levels to 62% and cellular ATP levels to 46% after 24 h. We conclude that pyocyanin is a key factor that redox cycles in the cytosol, generates H2O2, depletes glutathione and ATP, and impairs CFTR function in Pseudomonas-infected lungs.  相似文献   

11.
The reduction potential of a cell is related to its fate. Proliferating cells are more reduced than those that are differentiating, whereas apoptotic cells are generally the most oxidized. Glutathione is considered the most important cellular redox buffer and the average reduction potential (Eh) of a cell or organism can be calculated from the concentrations of glutathione (GSH) and glutathione disulfide (GSSG). In this study, triplicate groups of cod larvae at various stages of development (3 to 63 days post-hatch; dph) were sampled for analyses of GSSG/2GSH concentrations, together with activities of antioxidant enzymes and expression of genes encoding proteins involved in redox metabolism. The concentration of total GSH (GSH+GSSG) increased from 610±100 to 1260±150 μmol/kg between 7 and 14 dph and was then constant until 49 dph, after which it decreased to 810±100 μmol/kg by 63 dph. The 14- to 49-dph period, when total GSH concentrations were stable, coincides with the proposed period of metamorphosis in cod larvae. The concentration of GSSG comprised approximately 1% of the total GSH concentration and was stable throughout the sampling series. This resulted in a decreasing Eh from −239±1 to −262±7 mV between 7 and 14 dph, after which it remained constant until 63 dph. The changes in GSH and Eh were accompanied by changes in the expression of several genes involved in redox balance and signaling, as well as changes in activities of antioxidant enzymes, with the most dynamic responses occurring in the early phase of cod larval development. It is hypothesized that metamorphosis in cod larvae starts with the onset of mosaic hyperplasia in the skeletal muscle at approximately 20 dph (6.8 mm standard length (SL)) and ends with differentiation of the stomach and disappearance of the larval finfold at 40 to 50 dph (10–15 mm SL). Thus, metamorphosis in cod larvae seems to coincide with high and stable total concentrations of GSH.  相似文献   

12.
《Aquatic Botany》2007,86(4):353-360
The influence of cadmium (Cd) on physiological and biochemical parameters was studied to elucidate the mechanism of Cd resistance in Phragmites australis. Cadmium concentrations in roots, stems and leaves increased with exogenous Cd concentration, but Cd content in roots was much higher than in shoots. X-ray microanalysis was used to reveal compartments in which Cd accumulated in root cortex. Cadmium concentrations followed a gradient with the sequence: intercellular space > cell wall > vacuole > cytoplasm, indicating that most Cd was immobilized in the apoplast or sequestered into the vacuolar lumen. Sequential extraction of various Cd chelates revealed that more than half of extractable Cd was bound to proteins, whereas 26% was bound to organic acids. Cd-binding protein fractions were found in the roots after gel filtration chromatography, among which a polypeptide with an apparent molecular mass of 14 kDa bound Cd most avidly. One newly synthesized polypeptide of low molecular mass (1 kDa) appeared under Cd pollution, whereas a prominent fraction of 72 kDa disappeared. Four aldehyde oxidase (AO) isoenzyme activities increased significantly in roots under Cd pollution. Cd stress also enhanced xanthine dehydrogenase (XDH) activities in roots. Two AO polypeptides of different molecular sizes were detected in the roots by Western blot assay. The abundance of the 160 kDa subunit correlated with Cd stress, but the amount of the 90 kDa polypeptide did not change under Cd treatment. Enhanced abscisic acid (ABA) contents were observed in roots of P. australis exposed to Cd. The involvement of Cd distribution in plant tissues and subcellular compartments and of AO and XDH enzymatic activities in the acclimation mechanism of P. australis to Cd pollution is discussed herein.  相似文献   

13.
This work aimed to study the stability over time of the bacterial community in cæcum and fæces of the rabbit (diversity index and structure) without experimental disturbance and to evaluate its relationships with environmental parameters. Soft and hard fæces of 14 rabbits were sampled for 5 weeks while cæcal content was sampled on the 3rd week (by surgery) and the 5th week (at slaughter). Bacterial communities were assessed by studying CE-SSCP profiles of 16S rRNA genes fragments. Redox potential, pH, NH3-N concentration and volatile fatty acid concentrations were measured in the cæcum. Data showed that bacterial communities of soft and hard fæces barely differed from that of the cæcum (ANOSIM-R < 0.25; p < 0.05). Without disturbance, the bacterial communities of fæces were stable over time (ANOSIM-R < 0.25; p < 0.001). However, the bacterial communities of cæcum and fæces were affected by the surgery (ANOSIM-R = 0.22–0.33; p < 0.001). The cæcal content was an acidic (pH = 6.03 ± 0.33) and an anaerobic environment (redox potential = ?160 ± 43 mV). Only the redox potential was correlated with the diversity index of the bacterial community of the cæcum (R2 = 0.35; p < 0.05) and no environmental parameters were correlated to its structure.  相似文献   

14.
《Aquatic Botany》2007,87(3):235-241
Factors governing the germination of Chara vulgaris L. and Nitella flexilis L. were investigated in a series of three experiments. Temperature, light regimes, drying, and changes in oxidation–reduction potential are postulated to be possible triggers working in isolation or in combination on dormant oospores originating in sediments with different physical/chemical characteristics.In Experiment 1, none of the N. flexilis oospores inoculated into Petri plates and culture tubes containing either agar or water and then exposed to ‘normal’ 12 h daylight 12 h dark germinated when the redox of the media remained relatively constant over the 20-day observation period, while all of the oospores in a tube in which the redox of the agar declined to below 200 mV, germinated. Similarly no germination occurred in the dark without a decrease in redox. When redox fell to below 200 mV, 11% of the inoculated oospores germinated.In Experiment 2, N. flexilis and C. vulgaris oospores were subjected to cold pre-treatments and inoculated into either water or agar in both plates and tubes. Of the N. flexilis oospores 53% germinated in the dark, and 31% under normal lighting. Of the C. vulgaris, 3% germinated in the dark and 53% under normal lighting. The stimulation of germination with the decrease in redox in the media did not re-occur.Experiment 3 compared the germination of C. vulgaris oospores from four ponds contaminated by mining wastes and from a comparatively clean wetland. The oospores were again subjected (or not) to cold pre-treatments, and were drawn from both freshly concentrated sediments and air-dried sediments that had been held in storage for either 97 or 376 days. Notably, a consistent reduction in oospore viability occurred in oospores from a particular mine pond, contaminated by nickel tailings.  相似文献   

15.
Odor perception via the antennal sensilla in most honeybee species is poorly studied. We measured the antennal sensillum potential in response to Apis florea mandibular gland pheromone and showed that it is robust and reliable in forager and guard bees. Mandibular gland pheromone may be involved in signaling alarm or foraging resource depletion. Changes of antennal sensilla placodea potential of A. florea foragers and guards were measured after exposure to three concentrations of the synthetic pheromones, 2-heptanone and (Z)-11-eicosen-1-ol, using a potentiostat connected to an e-corder (ED401) with microelectrodes. The resting sensillum potential of A. florea foragers and guards were ?55.37 ± 3.44 and ?52.85 ± 5.34 mV, respectively. The sensillum potential of bees exposed to 1.0%, 5.0% and 10.0% (Z)-11-eicosen-1-ol were ?23.35 ± 0.98, ?16.78 ± 1.94 and ?24.24 ± 8.20 mV, respectively, in foragers, and ?21.95 ± 3.21, ?21.42 ± 4.73 and ?13.54 ± 4.16 mV, respectively, for guards. Exposure of bees to 1.0%, 5.0% and 10.0% 2-heptanone induced sensillum potentials of ?10.64 ± 2.44, ?44.88 ± 2.41 and ?48.84 ± 4.40 mV, respectively, in foragers and 15.85 ± 9.38, ?25.48 ± 1.43 and ?15.52 ± 6.61 mV, respectively, in guards. The highest sensillum potential was recorded in foragers exposed to 1.0% 2-heptanone. In general, except for the response to 1.0% 2-heptanone, the sensillum potentials of all bees to (Z)-11-eicosen-1-ol were higher than that of 2-heptanone. These results show that A. florea antennal sensilla in foragers and guard bees exhibit a stronger response to (Z)-11-eicosen-1-ol as compared to 2-heptanone. Our results also provide useful comparative data to explore olfactory perception in non-model honey bee species.  相似文献   

16.
Embryonic development involves dramatic changes in cell proliferation and differentiation that must be highly coordinated and tightly regulated. Cellular redox balance is critical for cell fate decisions, but it is susceptible to disruption by endogenous and exogenous sources of oxidative stress. The most abundant endogenous nonprotein antioxidant defense molecule is the tripeptide glutathione (γ-glutamylcysteinylglycine, GSH), but the ontogeny of GSH concentration and redox state during early life stages is poorly understood. Here, we describe the GSH redox dynamics during embryonic and early larval development (0–5 days postfertilization) in the zebrafish (Danio rerio), a model vertebrate embryo. We measured reduced and oxidized glutathione using HPLC and calculated the whole embryo total glutathione (GSHT) concentrations and redox potentials (Eh) over 0–120 h of zebrafish development (including mature oocytes, fertilization, midblastula transition, gastrulation, somitogenesis, pharyngula, prehatch embryos, and hatched eleutheroembryos). GSHT concentration doubled between 12 h postfertilization (hpf) and hatching. The GSH Eh increased, becoming more oxidizing during the first 12 h, and then oscillated around −190 mV through organogenesis, followed by a rapid change, associated with hatching, to a more negative (more reducing) Eh (−220 mV). After hatching, Eh stabilized and remained steady through 120 hpf. The dynamic changes in GSH redox status and concentration defined discrete windows of development: primary organogenesis, organ differentiation, and larval growth. We identified the set of zebrafish genes involved in the synthesis, utilization, and recycling of GSH, including several novel paralogs, and measured how expression of these genes changes during development. Ontogenic changes in the expression of GSH-related genes support the hypothesis that GSH redox state is tightly regulated early in development. This study provides a foundation for understanding the redox regulation of developmental signaling and investigating the effects of oxidative stress during embryogenesis.  相似文献   

17.
Coenzyme Q10 (Q10) is present in the circulation mainly in its reduced form (ubiquinol-10; UL10), but oxidizes quickly ex vivo to ubiquinone-10 (UN10). Therefore, native UL10:UN10 ratios, used as markers of redox status and disease risk, are difficult to measure. We established an RP-(U)HPLC method with coulometric detection to measure natively circulating UL10 and UN10 concentrations by adding a ubiquinol/ubiquinone mixture as an internal standard immediately after plasma preparation. This allowed adjustment for unavoidable artificial UL10 oxidation as well as for total losses (or gains) of analytes during sample storage, processing, and analysis because the internal standards exactly paralleled the chemical behavior of Q10. This technique applied to blood (n = 13) revealed Q10 levels of 680–3300 nM with a mean UL10:UN10 ratio of 95:5, which was inversely associated with total Q10 (r = ? 0.69; p = 0.004). The oxidation of UL10 to UN10 was equimolar, increased by O2, and decreased by lower temperatures or various degassing methods. Although UL10 was stable in blood or when pure in organic solvents at 22 °C, its oxidation was catalyzed dose dependently by α-tocopherol and butylated hydroxytoluene, particularly when present in combination. Key structural features for the catalytic pro-oxidant properties of phenolic antioxidants included two substituents vicinal to the phenolic hydroxyl group.  相似文献   

18.
Transition from fossil energy sources to biogas production has resulted in a strong increase of leakage accidents from fermenters, but knowledge on the effects of fermentation product runoff into freshwater systems is currently restricted to direct toxicity due to oxygen depletion. This study provides first information about the influence of digestate runoff on the physicochemical habitat properties and the bacterial community composition of the hyporheic interstitial which is important in determining ecosystem functioning. We exposed natural stream beds to different concentrations of two different digestates from fermenters (corn and manure feedstock), hypothesizing that the digestate addition causes acute changes of the physicochemical parameters and has distinct effects on microbial community composition of the hyporheic interstitial depending on concentration and type of digestate. In line with the hypotheses, pH value, conductivity, redox potential and ammonium differed significantly from controls and among treatments after digestate addition, but only for a maximum of two days. pH values (controls: 7.8; corn: 7.9; manure: 7.9) and conductivity (controls: 813 μS/cm; corn: 969 μS/cm; manure: 1097 μS/cm) increased, the redox potential (controls: 153 mV; corn: 145 mV; manure: 144 mV) decreased the first two days. A high peak of ammonium-N was detected in the corn and manure treatments (controls: 5 mg/l, corn: 80 mg/l; manure: 60 mg/l) at day 1. In contrast, changes in bacterial community composition were detectable for longer periods of time (>5 days). Seventeen unique T-RF fingerprints of bacterial community response to each of the different digestate treatments (11 unique T-RFs in manure and 6 unique T-RFs in corn treatments) were found, suggesting that this approach provides a suitable ecological indicator for source tracking, e.g. in case of a biogas power plant leakage accident.  相似文献   

19.
The role of honeybee mandibular gland compounds is poorly understood, although they may act as alarm pheromones. We measured forager and guard bee antennal responses evoked by two major components of mandibular gland secretions of the Asiatic honeybee, Apis cerana. Membrane potentials of antennal sensilla were measured after exposure to three concentrations of the synthetic alarm pheromones 2-heptanone and (Z)-11-eicosen-1-ol using a potentiostat (EA161) connected to an e-corder (ED401) with microelectrodes. The resting membrane potential of A. cerana foragers and guards was ?55.23 ± 1.44 and ?56.41 ± 1.21 mV, respectively. The membrane potential of foragers after exposure to 1.0, 5.0 and 10.0% 2-heptanone was ?5.32 ± 0.46, ?8.41 ± 1.33 and ?11.53 ± 2.16 mV, respectively. The membrane potential of guards was ?5.49 ± 1.66, ?8.46 ± 1.32 and ?7.31 ± 3.46 mV, respectively. Exposure of foragers to 1.0, 5.0 and 10.0% (Z)-11-eicosen-1-ol induced membrane potentials of ?24.00 ± 6.56, ?36.36 ± 5.18 and ?14.60 ± 8.20 mV, respectively; for guards they were ?47.62 ± 1.46, ?46.08 ± 0.87 and ?9.35 ± 1.96 mV, respectively. The highest membrane potential was found in foragers exposed to 1.0% 2-heptanone. The membrane potentials of foragers were higher than that of guards except at the highest concentration (10.0%) of both pheromones. These findings suggest that antennal sensory receptors of foragers may have higher specific thresholds than those of guards.  相似文献   

20.
The dioxygen reduction mechanism in cytochrome oxidases relies on proton control of the electron transfer events that drive the process. Proton delivery and proton channels in the protein that are relevant to substrate reduction and proton pumping are considered, and the current status of this area is summarized. We propose a mechanism in which the coupling of the oxygen reduction chemistry to proton translocation (P  F transition) is related to the properties of two groups of highly conserved residues, namely, His411/G386-T389 and the heme a3–propionateA–D399–H403 chain. This article is part of a Special Issue entitled: Respiratory Oxidases.  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