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.
The interactions of DNA-binding dyes (Hoechst 33258, DAPI, acridine orange) and DiBAC4(3) with hyperthermophilic archaeon Aeropyrum pernix cells were investigated by the combination of calorimetric, spectroscopic and microscopic techniques. All of the dyes, studied here, affect the thermal stability of DNA in vivo and in vitro. Hoechst 33258 is highly DNA-specific probe, which does not affect the thermal transitions of other cellular components as can be detected by differential scanning calorimetry (DSC). Due to this unique property, it can be used as a potential DNA marker for in vivo DSC studies. The localization of the dyes in the cells and viability assay was revealed by fluorescence microscopy. Hoechst 33258, DAPI and acridine orange did not distinguish between viable and non-viable cells of Aeropyrum pernix. Only with the commercially available Live/Dead BacLightTM kit we were able to discriminate viable and non-viable Aeropyrum pernix cells.  相似文献   

2.
Biomass of Desulfovibrio desulfuricans was used to recover Au(III) as Au(0) from test solutions and from waste electronic scrap leachate. Au(0) was precipitated extracellularly by a different mechanism from the biodeposition of Pd(0). The presence of Cu2+ (∼2000 mg/l) in the leachate inhibited the hydrogenase-mediated removal of Pd(II) but pre-palladisation of the cells in the absence of added Cu2+ facilitated removal of Pd(II) from the leachate and more than 95% of the Pd(II) was removed autocatalytically from a test solution supplemented with Cu(II) and Pd(II). Metal recovery was demonstrated in a gas-lift electrobioreactor with electrochemically generated hydrogen, followed by precipitation of recovered metal under gravity. A 3-stage bioseparation process for the recovery of Au(III), Pd(II) and Cu(II) is proposed.Victoria S. Baxter-Plant – Deceased  相似文献   

3.
In our study, we isolated the isolate Trichoderma SP2F1 from sediment samples from the Penchala River, heavily contaminated with effluents from nearby industrial areas. Qualitative and quantitative screening using plate and broth assay, respectively, supplemented with various concentrations of Cu(II) showed the isolate was able to tolerate 6 mM CuSO4, although growth was also detected in broths with 10 mM CuSO4. Trichoderma spp. was able to remove Cu(II) in aqueous solutions in both viable and non-viable cell forms. Bioaccumulation capacity of viable SP2F1 cells removed 19.60 mg g−1 of Cu(II) after 168 h incubation, while the maximum Cu(II) biosorption capacity for non-viable SP2F1 cells was 28.75 mg g−1 of Cu(II). Results here showed that Trichoderma spp isolate SP2F1 has good potential for application in Cu(II) removal, can be used to treat sewage waste by applying either in viable or non-viable cell forms.  相似文献   

4.
The surfaces of cells of Desulfovibrio desulfuricans,Desulfovibrio vulgaris and a new strain, Desulfovibrio sp. `Oz-7' were used to manufacturea novel bioinorganic catalyst via the reduction of Pd(II) to Pd(0) at the cell surface usinghydrogen as the electron donor. The ability of the palladium coated (palladised) cells to reductivelydehalogenate chlorophenol and polychlorinated biphenyl species was demonstrated. Dried, palladisedcells of D. desulfuricans, D. vulgaris and Desulfovibrio sp. `Oz-7'were more effective bioinorganic catalysts than Pd(II) reduced chemically under H2 orcommercially available finely divided Pd(0). Differences were observed in the catalyticactivity of the preparations when compared with each other. Negligible chloride release occurredfrom chlorophenol and polychlorinated biphenyls using biomass alone.  相似文献   

5.
The present study is the first report on the ability of Geobacter sulfurreducens PCA to reduce Pd(II) and produce Pd(0) nano-catalyst, using acetate as electron donor at neutral pH (7.0?±?0.1) and 30 °C. The microbial production of Pd(0) nanoparticles (NPs) was greatly enhanced by the presence of the redox mediator, anthraquinone-2,6-disulfonate (AQDS) when compared with controls lacking AQDS and cell-free controls. A cell dry weight (CDW) concentration of 800 mg/L provided a larger surface area for Pd(0) NPs deposition than a CDW concentration of 400 mg/L. Sample analysis by transmission electron microscopy revealed the formation of extracellular Pd(0) NPs ranging from 5 to 15 nm and X-ray diffraction confirmed the Pd(0) nature of the nano-catalyst produced. The present findings open the possibility for a new alternative to synthesize Pd(0) nano-catalyst and the potential application for microbial metal recovery from metal-containing waste streams.  相似文献   

6.
A novel bioinorganic catalyst was obtained via reduction of Pd(II) to Pd0 on to the surface of cells of Desulfovibrio desulfuricans at the expense of H2. Palladised biomass, supplied with formate or H2 as an electron donor, catalysed the dehalogenation of 2-chlorophenol and polychlorinated biphenyls. In the example of 2,3,4,5-tetrachlorobiphenyl, the bioinorganic catalyst promoted a rate of chloride release of 9.33 ± 0.17 nmol min–1 mg –1and only ~5% of this value was obtained using chemically reduced or commercially available Pd 0. In the case of 2,2,4,4,6,6-hexachlorobiphenyl the rate was more than four orders of magnitude faster than the degradation reported using a sulfidogenic culture. Negligible chloride release occurred from any of the chloroaromatic compounds using biomass alone, or from palladised biomass challenged with hexane carrier solvent only. Analysis of the spent solution showed that in addition to catalysis of reductive dehalogenation the new material was able to remove very effectively the organic residua, with neither any PCB nor any breakdown products identifiable by GC/MS.Revisions requested 8 September 2004; Revisions received 21 October 2004;  相似文献   

7.
Oxygen-evolving PS II particles from the thermophilic cyanobacterium Synechococcus elongatus are partially purified by centrifugation on a sucrose gradient and are bound to a Chelating Sepharose column loaded with Cu2+ ions. Bound particles are then transformed into PS II RC complexes by two washing steps. First, washing with a phosphate buffer (pH=6.5) containing 0.02% of SB 12 removes the rest of phycobilins and leaves pure PS II core particles on the column. Second, washing with a phosphate buffer (pH=6.2) containing 0.2 M LiClO4 and 0.05% of DM removes CP 47 and CP 43 and leaves bare PS II RC complexes on the column. These are then eluted with a phosphate buffer containing 1% of dodecylmaltoside (DM). The molar ratio of pigments in the eluate changes with the progress of elution but around the middle of the elution period a nearly stable ratio is maintained of Chl a: Pheo a: Car: Cyt b 559 equal to 2.9: 1: 0.9: 0.8. In these fractions the photochemical separation of charges could be demonstrated by accumulation of reduced pheophytin (A of 430–440 nm) and by the flash induced formation of P680+ (A at 820 nm). The relatively slow relaxation kinetics of the latter signal (t1/2 1 ms) may suggest that in a substantial fraction of the RCs QA remains bound to the complex.Abbreviations Car -carotene - Chl a chlorophyll a - CP43, CP47 chlorophyll-proteins, with Rm 43 and 47 kDa - DBMIB dibromothymoquinone,2,5-dibromo-3-methyl-6-isopropyl-1,4-benzoquinone - DM -dodecyl-d-maltoside - HPLC high-performance liquid chromatography - OG n-octyl--d-glucopyranoside - IMAC immobilied metal affinity chromatography - Pheo a pheophytin a - PQ-9 plastoquinone-9 - P680 primary electron donor in PS II - PS II RC Photosystem II reaction centre - QA primary electron acceptor in PS II - SB-12 N-dodecyl-N,N-dimethyl-3-amino-1-propanesulphonate, (sulphobetain 12)  相似文献   

8.
Leaf and stem explants of Cleome rosea formed calluses when cultured on MS medium supplemented with different concentrations of 2,4-dichlorophenoxyacetic acid (2,4-D) or 4-amino-3,5,6-trichloropicolinic acid (PIC). The highest biomass accumulation was obtained in the callus cultures initiated from stem explants on medium supplemented with 0.90 μM 2,4-D. Reddish-pink regions were observed on callus surface after 6–7 months in culture and these pigments were identified as anthocyanins. Anthocyanins production was enhanced by reducing temperature and increasing light irradiation. Pigmented calluses transferred to MS1/2 with a 1:4 ratio NH4+/NO3, 70 g L−1 sucrose and supplementation with 0.90 μM 2,4-D maintained a high biomass accumulation and showed an increase of 150% on anthocyanin production as compared with the initial culture conditions. Qualitative analysis of calluses was performed by high performance liquid chromatography coupled to diode array detector and electrospray ionization mass spectrometry (HPLC-DAD/ESIMS). Eleven anthocyanins were characterized and the majority of them were identified as acylated cyanidins, although two peonidins were also detected. The major peak was composed by two anthocyanins, whose proposed identity were cyanidin 3-(p-coumaroyl) diglucoside-5-glucoside and cyanidin 3-(feruloyl) diglucoside-5-glucoside.  相似文献   

9.
Industrial waste streams may contain contaminants that are valuable like Pd(II) and/or toxic and mutagenic like Cr(VI). Using Serratia sp. biofilm the former was biomineralized to produce a supported nanocrystalline Pd(0) catalyst, and this biofilm–Pd heterogeneous catalyst was then used to reduce Cr(VI) to less dangerous Cr(III) at room temperature, with formate as the electron donor. Cr(VI)(aq) is non‐paramagnetic while Cr(III)(aq) is paramagnetic, which enabled spatial mapping of Cr species concentrations within the reactor cell using non‐invasive magnetic resonance (MR) imaging experiments. Spatial reactivity heterogeneities were thus examined. In batch reactions, these could be attributed primarily to heterogeneity of Pd(0) distribution and to the development of gas bubbles within the reactor. In continuous flow reactions, spatial reactivity heterogeneities resulted primarily from heterogeneity of Cr(VI) delivery. Biotechnol. Bioeng. 2010;107: 11–20. © 2010 Wiley Periodicals, Inc.  相似文献   

10.
The increasing demand and limited natural resources for industrially important platinum‐group metal (PGM) catalysts render the recovery from secondary sources such as industrial waste economically interesting. In the process of palladium (Pd) recovery, microorganisms have revealed a strong potential. Hitherto, bacteria with the property of dissimilatory metal reduction have been in focus, although the biochemical reactions linking enzymatic Pd(II) reduction and Pd(0) deposition have not yet been identified. In this study we investigated Pd(II) reduction with formate as the electron donor in the presence of Gram‐negative bacteria with no documented capacity for reducing metals for energy production: Cupriavidus necator, Pseudomonas putida, and Paracoccus denitrificans. Only large and close‐packed Pd(0) aggregates were formed in cell‐free buffer solutions. Pd(II) reduction in the presence of bacteria resulted in smaller, well‐suspended Pd(0) particles that were associated with the cells (called “bioPd(0)” in the following). Nanosize Pd(0) particles (3–30 nm) were only observed in the presence of bacteria, and particles in this size range were located in the periplasmic space. Pd(0) nanoparticles were still deposited on autoclaved cells of C. necator that had no hydrogenase activity, suggesting a hydrogenase‐independent formation mechanism. The catalytic properties of Pd(0) and bioPd(0) were determined by the amount of hydrogen released in a reaction with hypophosphite. Generally, bioPd(0) demonstrated a lower level of activity than the Pd(0) control, possibly due to the inaccessibility of the Pd(0) fraction embedded in the cell envelope. Our results demonstrate the suitability of bacterial cells for the recovery of Pd(0), and formation and immobilization of Pd(0) nanoparticles inside the cell envelope. However, procedures to make periplasmic Pd(0) catalytically accessible need to be developed for future nanobiotechnological applications. Biotechnol. Bioeng. 2010;107: 206–215. © 2010 Wiley Periodicals, Inc.  相似文献   

11.
Microvillar cells (MCs) have been identified in the olfactory epithelium of various mammalian species from rodents to humans. Studies on properties and functions of MCs to date have yielded partially controversial results, supporting alternatively an epithelial or a neuronal nature of these cells. In the present study, single and double immunolabeling investigations were carried out using antibodies against cytoskeletal and integral membrane proteins in order to further characterize MCs in rat and mouse olfactory epithelium. Application of antibodies against ankyrin (ANK), a protein that links integral membrane proteins to the submembrane cytoskeleton, led to intense labeling of the basolateral membranes of numerous cells with characteristic MC morphology. ANK-immunoreactive (ir) cells bore an apical tuft of -actin-ir microvilli, were filled with cytokeratin 18 (CK18)-ir filamentous network, and extended a basal process that appeared to end above the basal membrane. Immunoreactions for villin, an actin-crosslinking protein particularly prominently expressed in brush cells in the gastrointestinal and respiratory tract epithelia, and for the -subunit of sodium-potassium ATPase (Na+, K+-ATPase), revealed that ANK-ir MCs fall into two subpopulations. The less frequent type I MCs displayed villin immunoreactivity in their apical microvilli and underneath the basolateral membranes; the more numerous type II MCs were negative for villin but possessed intense basolateral immunoreactivity for Na+, K+-ATPase. Strong reactivity for the epithelial-type integral membrane protein of adherens junctions, E-Cadherin, was localized in basolateral membranes of both types of MCs. Our results support an epithelial nature of ANK-ir MCs in rat and mouse olfactory epithelium. Type I MCs strongly resemble brush cells in their immunocytochemical characteristics, namely, their ANK reactivity, CK18 reactivity, and villin reactivity. The intense Na+, K+-ATPase reactivity of type II MCs implicates these cells in transport processes.  相似文献   

12.
Zhuang L  Zhou S  Wang Y  Liu Z  Xu R 《Bioresource technology》2011,102(7):4820-4826
This study demonstrated the feasibility to produce Bacillus thuringiensis subsp. kurstaki (Btk) based biopesticides using wastewater sludge as raw materials under solid-state fermentation (SSF). More than 1010 CFU/g viable cells of Btk were obtained using sludge or its mixture with agricultural wastes. This study well considered the effect of heavy metals on Btk growth and their changes of chemical speciation caused by SSF. The IC50 (concentration causing 50% inhibition in total cell biomass) for Pb(II), Cu(II), Cd(II) and Cr(III) on Btk were determined to be 227, 82, 15 and 263 mg/L, respectively. Exposure to 150 mg/L of Cu(II) severely reduced the amount and size of toxin crystals, which decreased the endotoxin synthesis and entomotoxicity potency of Btk cells. Using Tessier’s sequential extraction procedure, the exchangeable heavy metals in sludge were shown to be transformed into residual fractions after SSF, and thus significantly reduced their bioavailability and potential environmental risks.  相似文献   

13.
Microbial reduction of soluble Pd(II) by cells of Shewanella oneidensis MR-1 and of an autoaggregating mutant (COAG) resulted in precipitation of palladium Pd(0) nanoparticles on the cell wall and inside the periplasmic space (bioPd). As a result of biosorption and subsequent bioreduction of Pd(II) with H2, formate, lactate, pyruvate or ethanol as electron donors, recoveries higher than 90% of Pd associated with biomass could be obtained. The bioPd(0) nanoparticles thus obtained had the ability to reductively dehalogenate polychlorinated biphenyl (PCB) congeners in aqueous and sediment matrices. Bioreduction was observed in assays with concentrations up to 1000 mg Pd(II) l(-1) with depletion of soluble Pd(II) of 77.4% and higher. More than 90% decrease of PCB 21 (2,3,4-chloro biphenyl) coupled to formation of its dechlorination products PCB 5 (2,3-chloro biphenyl) and PCB 1 (2-chloro biphenyl) was obtained at a concentration of 1 mg l(-1) within 5 h at 28 degrees C. Bioreductive precipitation of bioPd by S. oneidensis cells mixed with sediment samples contaminated with a mixture of PCB congeners, resulted in dechlorination of both highly and lightly chlorinated PCB congeners adsorbed to the contaminated sediment matrix within 48 h at 28 degrees C. Fifty milligrams per litre of bioPd resulted in a catalytic activity that was comparable to 500 mg l(-1) commercial Pd(0) powder. The high reactivity of 50 mg l(-1) bioPd in the soil suspension was reflected in the reduction of the sum of seven most toxic PCBs to 27% of their initial concentration.  相似文献   

14.
Desulfovibrio desulfuricans reduces Pd(II) to Pd(0) at the expense of H2. Mass transfer limits the rate under hydrogen in a static solution, while a bubble reactor was inefficient due to loss of H2. A novel approach to the transfer of H2 to the biomass utilized a biofilm on the surface of a Pd-Ag membrane that traps and transports atomic hydrogen (H), formed at the back-side electrochemically, for delivery to the immobilized biofilm to form a biocatalytic surface for reduction of Pd(II) and deposition of Pd(0). Separation of the primary electrolysis chamber from the biocatalytic chamber permits the use of different solutions and pH in each, and use of a low voltage for H2 generation. Pd(0) recovery was efficient and fed by H2 on demand to give a clean, economic system with no generation of secondary wastes. The system was tested against a precious metal processing waste where the continuous removal of Pd, Pt and Rh was up to 88%, 99% and 75%, respectively, at a flow residence time of 10–20 min at an input pH of 2.5 and a total metals concentration of approx. 5 mM. Biorecovered Pd(0) was a better chemical catalyst than its chemical counterpart in a test reaction which liberated H2 from hypophosphite.  相似文献   

15.
A novel catalytic activity of palladium [Pd(0)]-coated cells of Desulfovibrio desulfuricans ATCC 29577 ["bio-Pd(0)"] is demonstrated. Reduction of 700 microM Cr(VI) occurred within 24 h using formate (25 mM) or hydrogen (1 atm) as the electron donor, under conditions whereby cells lacking bound Pd(0), or palladium metal manufactured via chemical reduction of soluble Pd(II), did not reduce Cr(VI). The biomass-bound Pd(0) also functioned in the continuous removal of 400 microM Cr(VI) from a 1 mM solution under H(2) (flow residence time approximately 5 h), where chemically prepared Pd(0) was ineffective. This demonstrates a new type of active bioinorganic catalysis, whereby the presence of biomass bound to Pd(0) confers a novel catalytic capability not seen with Pd base metal or biomass alone.  相似文献   

16.
Growth-decoupled cells of Desulfovibrio vulgaris NCIMB 8303 can be used to reduce Pd(II) to cell-bound Pd(0) (Bio-Pd0), a bioinorganic catalyst capable of reducing hexavalent chromium to less toxic Cr(III), using formate as the electron donor. Magnetic resonance imaging showed that Bio-Pd0, immobilized in chitosan and agar beads, is distinguishable from the surrounding gel and is evenly dispersed within the immobilization matrix. Agar-immobilized Bio-Pd0 and `chemical Pd0' were packed into continuous-flow reactors, and challenged with a solution containing 100 m Cr(VI) (pH 7) at a flow rate of 2.4 ml h–1. Agar-immobilized chemical Pd0 columns lost Cr(VI) reducing ability by 160 h, whereas columns containing immobilized Bio-Pd0 maintained 90% reduction until 680 h, after which reduction efficiency was gradually lost.  相似文献   

17.
Won SW  Park J  Mao J  Yun YS 《Bioresource technology》2011,102(4):3888-3893
A new type of biosorbent was developed for binding anionic precious metals through cross-linking waste biomass Corynebacterium glutamicum with polyethylenimine (PEI). This biomass was evaluated for the removal and recovery of palladium and compared to commercial adsorbents, such as Amberjet 4200 Cl, Lewatit Monoplus TP 214, SPC-100, and SPS-200. The kinetic experiments revealed that the sorption equilibrium was reached with 30 min for the PEI-modified biomass. The maximum uptake of the biosorbent was 176.8 mg/g, which was calculated using the Langmuir model. The Pd(II) maximum uptake exhibited the following order: Amberjet 4200 Cl > Lewatit Monoplus TP 214 > PEI-modified biomass > SPC-100 > SPS-200. Acidified thiourea in 1.0 M HCl was used to desorb Pd(II) from all of the sorbents examined.  相似文献   

18.
Cell-free supernatant from Shewanella oneidensis MR-1 reduced carbon tetrachloride to chloroform, a suspension of Fe(III) and solid Fe(III) to iron (II). The putative reducing agent was tentatively identified as menaquinone-1 (MQ-1)—a water-soluble menaquinone with a single isoprenoid residue in the side chain. Synthetic MQ-1 reduced carbon tetrachloride to chloroform and amorphous iron (III) hydroxide to iron (II). To test the generality of this result among menaquinones, the reductive activities of vitamin K2 (MQ-7)—a lipid-associated menaquinone with 7 or 8 isoprenoid residues—was evaluated. This molecule also reduced carbon tetrachloride to chloroform and iron (III) to iron (II). The results indicate that molecules within the menaquinone family may contribute to both the extracellular and cell-associated reduction of carbon tetrachloride and iron (III).  相似文献   

19.
A kinetic model that describes substrate interactions during reductive dehalogenation reactions is developed. This model describes how the concentrations of primary electron-donor and -acceptor substrates affect the rates of reductive dehalogenation reactions. A basic model, which considers only exogenous electron-donor and -acceptor substrates, illustrates the fundamental interactions that affect reductive dehalogenation reaction kinetics. Because this basic model cannot accurately describe important phenomena, such as reductive dehalogenation that occurs in the absence of exogenous electron donors, it is expanded to include an endogenous electron donor and additional electron acceptor reactions. This general model more accurately reflects the behavior that has been observed for reductive dehalogenation reactions. Under most conditions, primary electron-donor substrates stimulate the reductive dehalogenation rate, while primary electron acceptors reduce the reaction rate. The effects of primary substrates are incorporated into the kinetic parameters for a Monod-like rate expression. The apparent maximum rate of reductive dehalogenation (q m, ap ) and the apparent half-saturation concentration (K ap ) increase as the electron donor concentration increases. The electron-acceptor concentration does not affect q m, ap , but K ap is directly proportional to its concentration.Definitions for model parameters RX halogenated aliphatic substrate - E-M n reduced dehalogenase - E-M n+2 oxidized dehalogenase - [E-M n ] steady-state concentration of the reduced dehalogenase (moles of reduced dehalogenase per unit volume) - [E-M n+2] steady-state concentration of the oxidized dehalogenase (moles of reduced dehalogenase per unit volume) - DH2 primary exogenous electron-donor substrate - A primary exogenous electron-acceptor substrate - A2 second primary exogenous electron-acceptor substrate - X biomass concentration (biomass per unit volume) - f fraction of biomass that is comprised of the dehalogenase (moles of dehalogenase per unit biomass) - stoichiometric coefficient for the reductive dehalogenation reaction (moles of dehalogenase oxidized per mole of halogenated substrate reduced) - stoichiometric coefficient for oxidation of the primary electron donor (moles of dehalogenase reduced per mole of donor oxidized) - stoichiometric coefficient for oxidation of the endogenous electron donor (moles of dehalogenase reduced per unit biomass oxidized) - stoichiometric coefficient for reduction of the primary electron acceptor (moles of dehalogenase oxidized per mole of acceptor reduced) - stoichiometric coefficient for reduction of the second electron acceptor (moles of dehalogenase oxidized per mole of acceptor reduced) - r RX rate of the reductive dehalogenation reaction (moles of halogenated substrate reduced per unit volume per unit time) - r d1 rate of oxidation of the primary exogenous electron donor (moles of donor oxidized per unit volume per unit time) - r d2 rate of oxidation of the endogenous electron donor (biomass oxidized per unit volume per unit time) - r a1 rate of reduction of the primary exogenous electron acceptor (moles of acceptor reduced per unit volume per unit time) - r a2 rate of reduction of the second primary electron acceptor (moles of acceptor reduced per unit volume per unit time) - k RX mixed second-order rate coefficient for the reductive dehalogenation reaction (volume per mole dehalogenase per unit time) - k d1 mixed-second-order rate coefficient for oxidation of the primary electron donor (volume per mole dehalogenase per unit time) - k d2 mixed-second-order rate coefficient for oxidation of the endogenous electron donor (volume per mole dehalogenase per unit time) - b first-order biomass decay coefficient (biomass oxidized per unit biomass per unit time) - k a1 mixed-second-order rate coefficient for reduction of the primary electron acceptor (volume per mole dehalogenase per unit time) - k a2 mixed-second-order rate coefficient for reduction of the second primary electron acceptor (volume per mole dehalogenase per unit time) - q m,ap apparent maximum specific rate of reductive dehalogenation (moles of RX per unit biomass per unit time) - K ap apparent half-saturation concentration for the halogenated aliphatic substrate (moles of RX per unit volume) - k ap apparent pseudo-first-order rate coefficient for reductive dehalogenation (volume per unit biomass per unit time)  相似文献   

20.
The mechanism of nitrate transport across the tonoplast of barley root cells   总被引:14,自引:0,他引:14  
Nitrate-selective microelectrodes were used to measure not only nitrate activity in the cytoplasm and vacuole of barley (Hordeum vulgare L.) root cells, but also the tonoplast electrical membrane potential. For epidermal cells, the mean cytoplasmic and vacuolar pNO3 (-log10 [NO3]) values were 2.3±0.04 (n=19) and 1.41±0.03 (n=35), respectively, while for cortical cells, the mean cytoplasmic and vacuolar nitrate values were 2.58±0.18 (n=4) and 1.17±0.06 (n=13), respectively. These results indicate that the accumulation of nitrate in the vacuole must be an active process. Proton-selective microelectrodes were used to measure the proton gradient across the tonoplast to assess the possibility that nitrate transport into the vacuole is mediated by an H+/NO 3 antiport mechanism. For epidermal cells, the mean cytoplasmic and vacuolar pH values were 7.12±0.06 (n=10) and 4.93±0.11 (n=22), respectively, while for cortical cells, the mean cytoplasmic and vacuolar pH values were 7.24±0.07 (n=3) and 5.09±0.17 (n=7), respectively. Calculations of the energetics for this mechanism indicate that the observed gradient of nitrate across the tonoplast of both epidermal and cortical cells could be achieved by an H+/NO 3 antiport with a 11 stoichiometry.Abbreviations and Symbols G/F free-energy change for H+/NO 3 antiport - F Faraday constant - pHc cytoplasmic pH - pHv vacuolar pH - p[NO3]c log10 (cytoplasmic [NO 3 ]) - P[NO3]v -log10 (vacuolar [NO3]) We wish to thank Dr. K. Moore for assistance with statistical analysis.  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